14.07.2014 Views

Stochastic Volatility and Seasonality in ... - Interconti, Limited

Stochastic Volatility and Seasonality in ... - Interconti, Limited

Stochastic Volatility and Seasonality in ... - Interconti, Limited

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Stochastic</strong> <strong>Volatility</strong> <strong>and</strong> <strong>Seasonality</strong> <strong>in</strong> Commodity<br />

Futures <strong>and</strong> Options: The Case of Soybeans<br />

Mart<strong>in</strong> Christian Richter <strong>and</strong> Carsten Sørensen<br />

Department of F<strong>in</strong>ance<br />

Copenhagen Bus<strong>in</strong>ess School<br />

Solbjerg Plads 3<br />

DK-2000 Frederiksberg<br />

Denmark


<strong>Stochastic</strong> <strong>Volatility</strong> <strong>and</strong> <strong>Seasonality</strong> <strong>in</strong> Commodity<br />

Futures <strong>and</strong> Options: The Case of Soybeans<br />

Abstract<br />

We estimate a cont<strong>in</strong>uous-time stochastic volatility model us<strong>in</strong>g panel data of soybean<br />

futures <strong>and</strong> options on soybean futures. The model of commodity price dynamics<br />

is with<strong>in</strong> the class of so-called aff<strong>in</strong>e asset pric<strong>in</strong>g models, <strong>and</strong> option prices<br />

are determ<strong>in</strong>ed us<strong>in</strong>g a st<strong>and</strong>ard <strong>in</strong>version of characteristic functions approach. Our<br />

model<strong>in</strong>g acknowledges that commodities exhibit seasonality patterns <strong>in</strong> both spot<br />

price level <strong>and</strong> volatility <strong>and</strong>, hence, the <strong>in</strong>volved commodity dynamics <strong>in</strong>volve<br />

stochastic differential equations that are <strong>in</strong>homogeneous <strong>in</strong> time. The estimation<br />

method is based on a state space formulation of the model <strong>and</strong> a quasi maximum<br />

likelihood approach. Estimation results are obta<strong>in</strong>ed based on weekly observations<br />

of soybean futures prices <strong>and</strong> options prices from the Chicago Board of Trade <strong>in</strong> the<br />

period October 1984 to March 1999. Our empirical results support the conceptual<br />

ideas <strong>in</strong> the theory of storage, though not the view that convenience yields are like<br />

tim<strong>in</strong>g options.<br />

JEL Classification: G13, C00


1 Introduction<br />

This paper sets up <strong>and</strong> estimates a cont<strong>in</strong>uous-time model of commodity price behavior us<strong>in</strong>g<br />

panel data of soybean futures prices <strong>and</strong> soybean option prices us<strong>in</strong>g data from the Chicago<br />

Board of Trade (CBOT) over the period 1984-1999. The approach is conceptually related<br />

to the analysis <strong>in</strong> Schwartz (1997), who sets up <strong>and</strong> estimates models of commodity futures<br />

prices us<strong>in</strong>g the Kalman-filter <strong>and</strong> a state-space approach. 1 However, the model <strong>in</strong> this paper<br />

<strong>in</strong>cludes additional realistic features s<strong>in</strong>ce we, <strong>in</strong> particular, allow for stochastic volatility<br />

<strong>and</strong> seasonality <strong>in</strong> the theoretical model<strong>in</strong>g of commodity price behavior. These features are<br />

especially important for agricultural commodities with a harvest<strong>in</strong>g pattern that <strong>in</strong>duces a<br />

seasonal component <strong>in</strong> the commodity price as well as <strong>in</strong> the volatility of the commodity price.<br />

The seasonality feature <strong>in</strong> both commodity prices <strong>and</strong> volatilities is thus known to be one of the<br />

empirical characteristics that make commodities strik<strong>in</strong>gly different from stocks, bonds, <strong>and</strong><br />

other conventional f<strong>in</strong>ancial assets; see e.g. the discussion <strong>in</strong> Routledge, Seppi <strong>and</strong> Spatt (2000).<br />

Furthermore, as an <strong>in</strong>novative feature, our panel data estimation approach is based on both<br />

futures prices as well as option prices across different exercise prices <strong>and</strong> maturities.<br />

The commodity price behavior is modeled <strong>in</strong> cont<strong>in</strong>uous-time as part of a system of stochastic<br />

differential equations with<strong>in</strong> the class of asset pric<strong>in</strong>g models where the drift <strong>and</strong> diffusion<br />

coefficients are aff<strong>in</strong>e functions of the relevant basic state-variables. In our framework there are<br />

three basic state-variables: the commodity spot price, the convenience yield, <strong>and</strong> the volatility<br />

of the spot price. Aff<strong>in</strong>e models are tractable for asset pric<strong>in</strong>g purposes <strong>and</strong>, <strong>in</strong> our case,<br />

for the valuation of commodity futures contracts <strong>and</strong> options written on commodity futures.<br />

Due to the tractability of aff<strong>in</strong>e models, they are widely used for model<strong>in</strong>g of term structures<br />

of <strong>in</strong>terest rates <strong>and</strong> pric<strong>in</strong>g options on equity <strong>and</strong> foreign exchange rates. 2 In particular, it<br />

is <strong>in</strong> general possible to determ<strong>in</strong>e relevant conditional characteristic functions with<strong>in</strong> aff<strong>in</strong>e<br />

models which <strong>in</strong> turn facilitates the evaluation of option prices through an <strong>in</strong>version approach,<br />

build<strong>in</strong>g on the ideas suggested by Heston (1993) for a specific stochastic volatility model of<br />

1 See also Schwartz <strong>and</strong> Smith (2000) <strong>and</strong> Sørensen (2001) for empirical analyzes of commodity price models<br />

follow<strong>in</strong>g this approach.<br />

2 See e.g. Duffie <strong>and</strong> Kan (1996) for a general specification of aff<strong>in</strong>e term structure models, Heston (1993)<br />

<strong>and</strong> Bakshi, Cao <strong>and</strong> Chen (1997) for aff<strong>in</strong>e models of equity returns, <strong>and</strong> Backus, Foresi <strong>and</strong> Telmer (1996) <strong>and</strong><br />

Bates (1996) for aff<strong>in</strong>e models of foreign exchange rates. Duffie, Pan <strong>and</strong> S<strong>in</strong>gleton (2000) provide a discussion<br />

of other applications as well as a general treatment of the aff<strong>in</strong>e class of models, with the possibility of jumps<br />

<strong>in</strong>cluded.<br />

1


currency <strong>and</strong> equity returns. The model<strong>in</strong>g of commodity spot price behavior <strong>in</strong> this paper<br />

differs from models of <strong>in</strong>terest rates, equity returns, <strong>and</strong> foreign exchange by the fact that<br />

the <strong>in</strong>volved stochastic differential equations are <strong>in</strong>homogeneous <strong>in</strong> time s<strong>in</strong>ce the drift <strong>and</strong><br />

diffusion coefficients are functions of calender time, due to the seasonality feature, <strong>and</strong> by<br />

the <strong>in</strong>clusion of a convenience yield. On the other h<strong>and</strong>, we build on the same traditional<br />

no-arbitrage approach to the pric<strong>in</strong>g of commodity derivatives as used <strong>in</strong> analyz<strong>in</strong>g the term<br />

structure of <strong>in</strong>terest, equity derivatives, <strong>and</strong> foreign exchange derivative, <strong>and</strong> also used widely<br />

for analyz<strong>in</strong>g commodity contracts <strong>and</strong> commodity hedg<strong>in</strong>g <strong>in</strong> related papers. 3<br />

The estimation approach is based on quasi maximum likelihood <strong>and</strong> the panel data estimation<br />

approach suggested orig<strong>in</strong>ally by Chen <strong>and</strong> Scott (1993) <strong>in</strong> the context of models of<br />

the term structure of <strong>in</strong>terest rates. In particular, at each sample date we observe a panel<br />

of futures prices <strong>and</strong> option prices which are related to the basic state-variables through a<br />

non-l<strong>in</strong>ear measurement equation. The measurement equation is based on the pric<strong>in</strong>g results<br />

that we establish for the specific aff<strong>in</strong>e model of commodity price behavior with respect to<br />

relevant futures <strong>and</strong> option prices. Follow<strong>in</strong>g the ideas <strong>in</strong> Chen <strong>and</strong> Scott (1993), we assume<br />

that all but three contracts are observed with measurement error. The latent value of the<br />

basic state-variables can be exactly filtered out at each sample date by <strong>in</strong>version based on the<br />

three contract prices which are observed without error, <strong>and</strong> the likelihood of the full sample<br />

can then be described based on the likelihood of the filtered paths of the state-variables. In the<br />

evaluation of the likelihood of the filtered state-variables we apply a quasi maximum likelihood<br />

approach where the relevant densities are substituted by Gaussian densities with appropriate<br />

first- <strong>and</strong> second order moments. The estimation approach is analogues to the estimation<br />

approach used <strong>in</strong> related papers on <strong>in</strong>terest rates <strong>and</strong> equity returns <strong>in</strong> aff<strong>in</strong>e asset pric<strong>in</strong>g<br />

models. 4<br />

3 See e.g. Black (1976), Brennan <strong>and</strong> Schwartz (1985), McDonald <strong>and</strong> Siegel (1986), Gibson <strong>and</strong><br />

Schwartz (1990), Jamshidian <strong>and</strong> Fe<strong>in</strong> (1990), Bjerksund (1991), Cortazar <strong>and</strong> Schwartz (1994), Dixit <strong>and</strong><br />

P<strong>in</strong>dyck (1994), Schwartz (1997), Hilliard <strong>and</strong> Reis (1998), Miltersen <strong>and</strong> Schwartz (1998), <strong>and</strong> Schwartz <strong>and</strong><br />

Smith (2000) for advances on no-arbitrage models with a focus on the valuation of commodity cont<strong>in</strong>gent claims,<br />

such as real options, <strong>and</strong> hedg<strong>in</strong>g.<br />

4 See e.g. Chen <strong>and</strong> Scott (1993) <strong>and</strong> Dai <strong>and</strong> S<strong>in</strong>gleton (2000) for estimation of models of the term structure of<br />

<strong>in</strong>terest rates, Duffie <strong>and</strong> S<strong>in</strong>gleton (1997) for estimation of a model of swap yields, Duffie <strong>and</strong> S<strong>in</strong>gleton (1999)<br />

<strong>and</strong> Duffee (1999) for models of credit risky bonds, <strong>and</strong> Bates (2000) <strong>and</strong> Chernov <strong>and</strong> Ghyssel (2000) for<br />

estimations of equity models us<strong>in</strong>g options data. This list of related papers concerned with estimation of aff<strong>in</strong>e<br />

asset pric<strong>in</strong>g models is very partial, <strong>and</strong> this is a fast grow<strong>in</strong>g research area. Additional relevant references on<br />

2


The estimated commodity price model describes the simultaneous dynamics of the commodity<br />

spot price, the convenience yield, <strong>and</strong> the volatility of the commodity spot price. When<br />

discuss<strong>in</strong>g our specific estimation results, based on soybean futures <strong>and</strong> option contracts from<br />

CBOT, we are <strong>in</strong> particular <strong>in</strong>terested <strong>in</strong> how the results relate to the basic ideas of the theory<br />

of storage by Kaldor (1939), Work<strong>in</strong>g (1948, 1949), Brennan (1958), <strong>and</strong> Telser (1958). The<br />

theory of storage expla<strong>in</strong>s the differences between the spot <strong>and</strong> futures prices with different<br />

contract horizons <strong>in</strong> terms of advantages of physical <strong>in</strong>ventory compared to advantages of<br />

ownership of a futures contract for future delivery of the commodity. Physical <strong>in</strong>ventory is<br />

associated with cost of carry such as storage costs <strong>and</strong> the <strong>in</strong>terest forgone by <strong>in</strong>vest<strong>in</strong>g <strong>in</strong><br />

storage. On the other h<strong>and</strong>, physical <strong>in</strong>ventory gives rise to a convenience yield from be<strong>in</strong>g<br />

able to profit from temporary price <strong>in</strong>creases due to temporary shortages of the particular<br />

commodity or from be<strong>in</strong>g able to ma<strong>in</strong>ta<strong>in</strong> a production process despite abrupt shortages of<br />

the commodity used as <strong>in</strong>put. 5 The theory of storage predicts a negative relationship between<br />

supply/<strong>in</strong>ventories <strong>and</strong> convenience yields. Thus, when <strong>in</strong>ventories are full <strong>and</strong> supply is plenty<br />

the convenience yield from marg<strong>in</strong>al storage is low, <strong>and</strong> vice versa when <strong>in</strong>ventories are close<br />

to be<strong>in</strong>g empty. 6 Furthermore, s<strong>in</strong>ce hold<strong>in</strong>g <strong>in</strong>ventory allows for flexibility <strong>in</strong> the tim<strong>in</strong>g of<br />

the use of a given commodity, the convenience yield may be viewed as a tim<strong>in</strong>g option. Recently,<br />

Routledge, Seppi <strong>and</strong> Spatt (2000) has thus provided a formal model support<strong>in</strong>g this<br />

idea build<strong>in</strong>g on the apparatus of Wright <strong>and</strong> Williams (1989), Williams <strong>and</strong> Wright (1991),<br />

Deaton <strong>and</strong> Laroque (1992, 1996), Chambers <strong>and</strong> Bailey (1996), among others, who formalize<br />

the theory of storage <strong>in</strong> competitive rational expectations models.<br />

Our empirical results are to some degree <strong>in</strong> accordance with the theory of storage. For<br />

example, the seasonal components <strong>in</strong> both the soybean convenience yields <strong>and</strong> volatilities are<br />

at the highest just before harvest<strong>in</strong>g when <strong>in</strong>ventories are low <strong>and</strong> supply scarce. The spot<br />

price is estimated to be positively correlated with the convenience yield, so this is also a period<br />

where spot prices are relatively high. While this supports the ideas <strong>in</strong> the theory of storage,<br />

related literature can be found <strong>in</strong> e.g. S<strong>in</strong>gleton (2000) <strong>and</strong> Pan (2001).<br />

5 When we refer to the convenience yield <strong>in</strong> the follow<strong>in</strong>g <strong>and</strong> <strong>in</strong> the formal cont<strong>in</strong>uous-time model, we are<br />

<strong>in</strong> fact referr<strong>in</strong>g to the net convenience yield, i.e. the convenience yield net of storage costs, unless otherwise<br />

specified.<br />

6 S<strong>in</strong>ce <strong>in</strong>ventories/supply of agricultural commodities vary with the season, a seasonality pattern <strong>in</strong> convenience<br />

yields is, therefore, well <strong>in</strong> accordance with the prediction of the theory of storage, as discussed by e.g.<br />

Fama <strong>and</strong> French (1987).<br />

3


our results on the other h<strong>and</strong> do not support the view that the convenience yield behaves<br />

like a tim<strong>in</strong>g option. In particular, our estimation results suggest that convenience yields <strong>and</strong><br />

volatilities are slightly negatively correlated contrary to how an option depends on volatility.<br />

Furthermore, these estimation results are consistent across different time periods.<br />

Furthermore, our empirical results demonstrate that the soybean commodity spot price is<br />

positively correlated with volatility. This is opposed to the so-called leverage effect <strong>in</strong> equity<br />

markets where the correlation between stocks <strong>and</strong> volatility is usually estimated negative. This<br />

estimation result is, however, consistent with the observation that the implied volatility “smile”<br />

or “smirk” <strong>in</strong> soybean options has opposite slope of the post-1987 crash “smirk” observed for<br />

equity options.<br />

The paper is organized as follows. In section 2, the model <strong>and</strong> estimation approach is<br />

described <strong>in</strong> details <strong>and</strong> <strong>in</strong> section 3 the data <strong>and</strong> estimation results are presented. A f<strong>in</strong>al<br />

section concludes.<br />

2 The model <strong>and</strong> estimation approach<br />

Let W = (W 1 , W 2 , W 3 ) be a three-dimensional Brownian motion def<strong>in</strong>ed on a probability space<br />

(Ω, F, P). F = {F t : t ≥ 0} denotes the st<strong>and</strong>ard filtration of W <strong>and</strong>, formally, (Ω, F, F, P)<br />

is the basic model for uncerta<strong>in</strong>ty <strong>and</strong> <strong>in</strong>formation arrival <strong>in</strong> the follow<strong>in</strong>g. We will assume<br />

that W has a constant “<strong>in</strong>stantaneous” correlation matrix Σ with entries ρ ij <strong>and</strong> ones <strong>in</strong> the<br />

diagonal.<br />

2.1 Commodity price dynamics<br />

The model of commodity price behavior <strong>in</strong>corporates important realistic features such as<br />

stochastic convenience yields <strong>and</strong> stochastic volatility as well as a seasonal feature. Let P t<br />

denote the spot commodity price at time t. The convenience yield <strong>and</strong> the seasonal adjusted<br />

spot price volatility at time t are denoted δ t <strong>and</strong> v t , respectively. The dynamics of the threedimensional<br />

process (P, δ, v) are described by the follow<strong>in</strong>g stochastic differential equation<br />

system,<br />

dP t = P t<br />

[<br />

(r − δ t + λ P e ν(t) v t ) dt + e ν(t)√ v t dW 1,t<br />

]<br />

dδ t = (α(t) − βδ t + λ δ σ δ e ν(t) v t ) dt + σ δ e ν(t)√ v t dW 2,t (2)<br />

dv t = (θ − κv t + λ v σ v v t ) dt + σ v<br />

√<br />

vt dW 3,t (3)<br />

(1)<br />

4


where β, κ, θ, λ P , λ δ , λ v , σ δ , <strong>and</strong> σ v , as well as the correlation parameters of W , are constant<br />

parameters that we will estimate <strong>in</strong> the subsequent sections. α(t) <strong>and</strong> ν(t) are determ<strong>in</strong>istic<br />

functions that aim at describ<strong>in</strong>g a seasonal pattern <strong>in</strong> convenience yields <strong>and</strong> volatilities; the<br />

functional forms of α(t) <strong>and</strong> ν(t) are described <strong>and</strong> discussed below.<br />

The parameters λ P , λ δ , <strong>and</strong> λ v are the risk premia on commodity price uncerta<strong>in</strong>ty, convenience<br />

yield uncerta<strong>in</strong>ty, <strong>and</strong> volatility uncerta<strong>in</strong>ty. We will assume markets are complete <strong>and</strong><br />

the existence of an equivalent mart<strong>in</strong>gale measure, Q, so that prices on cont<strong>in</strong>gent claims can<br />

be uniquely derived by evaluat<strong>in</strong>g the relevant expectations of the discounted payoff under Q.<br />

Us<strong>in</strong>g a st<strong>and</strong>ard term<strong>in</strong>ology, we will also refer to Q as the risk-neutral probability measure<br />

<strong>in</strong> the follow<strong>in</strong>g. Under the risk-neutral probability measure, risk premia are zero <strong>and</strong> the<br />

dynamics of (P, δ, v) are described by,<br />

dP t = P t<br />

[(r − δ t ) dt + e ν(t)√ v t dW Q 1,t<br />

dδ t = (α(t) − βδ t ) dt + e ν(t) √<br />

σ δ vt dW Q 2,t (5)<br />

√<br />

dv t = (θ − κv t ) dt + σ v vt dW Q 3,t (6)<br />

From equations (1) <strong>and</strong> (4) it follows that the convenience yield, δ t , can be <strong>in</strong>terpreted as<br />

the return shortfall on an <strong>in</strong>ventory position of the specific commodity. 7 Basically, <strong>in</strong> order<br />

to break-even on the marg<strong>in</strong>al <strong>in</strong>ventory position, there must be a ga<strong>in</strong>, or convenience yield,<br />

from hold<strong>in</strong>g <strong>in</strong>ventory to exactly offset the return shortfall.<br />

The parameters β <strong>and</strong> κ determ<strong>in</strong>e the degree of reversion to the determ<strong>in</strong>istic seasonal<br />

pattern <strong>in</strong> the convenience yield <strong>and</strong> the long-run volatility level θ, respectively. Likewise,<br />

σ δ <strong>and</strong> σ v are parameters that, together with the level of v t , determ<strong>in</strong>e the volatility of the<br />

convenience yield <strong>and</strong> the volatility of the stochastic volatility of the commodity price. The<br />

parameter r is the short default-free <strong>in</strong>terest rate which is assumed constant (but we allow to<br />

vary over time <strong>in</strong> our empirical analysis). 8<br />

The determ<strong>in</strong>istic seasonal functions α(t) <strong>and</strong> ν(t) are def<strong>in</strong>ed as,<br />

∑<br />

α(t) = α 0 + (α k cos (2πkt) + αk ∗ s<strong>in</strong> (2πkt)) (7)<br />

K α<br />

k=1<br />

7 In fact, the convenience yield may be def<strong>in</strong>ed as the return shortfall on hold<strong>in</strong>g the commodity; see, e.g.,<br />

MacDonald <strong>and</strong> Siegel (1984).<br />

8 Our pric<strong>in</strong>g results will carry through if <strong>in</strong>terest rates are stochastic but distributed <strong>in</strong>dependently of the<br />

soybean dynamics. In this case, discount<strong>in</strong>g is done us<strong>in</strong>g the appropriate zero coupon bonds, as <strong>in</strong> our empirical<br />

approach.<br />

]<br />

(4)<br />

5


<strong>and</strong><br />

∑K ν<br />

ν(t) = (ν k cos (2πkt) + νk ∗ s<strong>in</strong> (2πkt)) (8)<br />

k=1<br />

where K α <strong>and</strong> K ν determ<strong>in</strong>e the number of terms <strong>in</strong> the sums <strong>and</strong> α 0 , α k , αk ∗, k = 1, . . . , Kα ,<br />

<strong>and</strong> ν k , νk ∗, k = 1, . . . , Kν are constant parameters to be estimated. The specific functional<br />

form of the determ<strong>in</strong>istic seasonal components were orig<strong>in</strong>ally suggested by Hannan, Terrell,<br />

<strong>and</strong> Tuckwell (1970) as an alternative to st<strong>and</strong>ard dummy variable methods <strong>in</strong> econometric<br />

model<strong>in</strong>g of seasonality. Due to the cont<strong>in</strong>uous time feature of our analysis this is a flexible<br />

<strong>and</strong> natural choice of model<strong>in</strong>g the seasonal aspects of commodity price behavior which is also<br />

applied <strong>in</strong> Sørensen (2001).<br />

2.2 Futures prices<br />

Follow<strong>in</strong>g Cox, Ingersoll <strong>and</strong> Ross (1981), the futures price at time t on a futures contract<br />

for delivery at time T ≥ t can be obta<strong>in</strong>ed by tak<strong>in</strong>g the relevant expectations of the future<br />

spot price, F t (T ) = E Q t [P T ]. Equivalently, by us<strong>in</strong>g the Feynman-Kac formula, F t (T ) =<br />

F (P t , δ t , v t , t; T ) where F (·) can be found as the solution to a partial differential equation on<br />

the form<br />

1<br />

2 e2ν(t) vP 2 ∂2 F<br />

+ 1 ∂P 2 2 σ2 δ e2ν(t) v ∂2 F<br />

∂δ 2<br />

+ ρ 23 σ δ σ v e ν(t) v ∂2 F<br />

∂F<br />

∂δ∂v<br />

+ (r − δ)P<br />

∂P<br />

+ 1 2 σ2 vv ∂2 F<br />

∂v 2<br />

+ ρ 12 σ δ e 2ν(t) vP ∂2 F<br />

∂P ∂δ + ρ 13σ v e ν(t) vP ∂2 F<br />

∂P ∂v<br />

+ (α(t) − βδ)<br />

∂F<br />

∂δ<br />

+ (θ − κv)<br />

∂F<br />

∂v + ∂F<br />

∂t<br />

= 0<br />

(9)<br />

with term<strong>in</strong>al condition F (P, δ, v, T ; T ) = P .<br />

For a given <strong>and</strong> fixed expiration date T , the solution to the partial differential equation <strong>in</strong><br />

(9) is given by<br />

F (P, δ, v, t; T ) = P e A(t;T )+B(t;T )v+D(t;T )δ (10)<br />

where<br />

D(t; T ) = − 1 β<br />

(<br />

1 − e −β(T −t)) (11)<br />

<strong>and</strong> A(t; T ) <strong>and</strong> B(t; T ) are the solutions to the ord<strong>in</strong>ary differential equations<br />

1<br />

2 σ2 δ e2ν(t) (D(t; T )) 2 + ρ 12 σ δ e 2ν(t) D(t; T ) + 1 2 σ2 v (B(t; T )) 2<br />

+ (ρ 13 σ v e ν(t) + ρ 23 σ δ σ v e ν(t) D(t; T ) − κ)B(t; T ) + B ′ (t; T ) = 0 , B(T ; T ) = 0<br />

(12)<br />

<strong>and</strong><br />

r + α(t)D(t; T ) + θB(t; T ) + A ′ (t; T ) = 0 , A(T ; T ) = 0 (13)<br />

6


Solutions to the ord<strong>in</strong>ary differential equations <strong>in</strong> (11) <strong>and</strong> (13) are not available <strong>in</strong> closed form.<br />

Solutions are, however, easily obta<strong>in</strong>ed numerically with very high precision <strong>and</strong> speed us<strong>in</strong>g<br />

for example Runge-Kutta methods; this is the approach followed <strong>in</strong> our empirical analysis.<br />

2.3 Option prices<br />

We will consider options written on commodity futures. By us<strong>in</strong>g Ito’s lemma on the futures<br />

price, as described <strong>in</strong> equation (10), it can be seen that the dynamics of the futures price under<br />

the equivalent mart<strong>in</strong>gale measure, Q, are given by<br />

dF t (T ) = F t<br />

√<br />

vt<br />

[<br />

e ν(t) dW Q 1,t + σ vB(t; T ) dW Q 3,t + σ δe ν(t) D(t; T ) dW Q 2,t . ]<br />

Let C t be the price of a European call option written of a the futures price of a futures contract<br />

expir<strong>in</strong>g at time T . The option has strike price K <strong>and</strong> matures at time τ; t ≤ τ ≤ T . The call<br />

price can be determ<strong>in</strong>ed by tak<strong>in</strong>g the relevant expectations,<br />

C t = E Q t<br />

(14)<br />

[<br />

]<br />

e −r(τ−t) max[0, F τ (T ) − K] . (15)<br />

From the Feynman-Kac’s formula, <strong>and</strong> the dynamics of F t <strong>and</strong> v t <strong>in</strong> equations (14) <strong>and</strong> (6), it<br />

follows that the option price can be represented as C t = C(F t , v t , t) where the function C(·)<br />

is the solution to the partial differential equation<br />

where<br />

1<br />

2 σ2 F (t; T )vF 2 ∂2 C<br />

+ 1 ∂F 2 2 σ2 vv ∂2 C<br />

+ σ<br />

∂v 2 F v (t; T )vF ∂2 C<br />

∂C<br />

∂F ∂v<br />

+ (θ − κv)<br />

∂v + ∂C<br />

∂t − rC = 0 (16)<br />

σ 2 F (t; T ) = e2ν(t) + σ 2 v (B(t; T )) 2 + σ 2 δ e2ν(t) (D(t; T )) 2 + 2ρ 13 σ v e ν(t) B(t; T )<br />

+ 2ρ 12 σ δ e 2ν(t) D(t; T ) + 2ρ 23 σ v σ δ e ν(t) B(t; T )D(t; T )<br />

<strong>and</strong><br />

σ F v (t; T ) = ρ 13 σ v e ν(t) + σ 2 vB(t; T ) + ρ 23 σ v σ δ e ν(t) D(t; T )<br />

<strong>and</strong> with term<strong>in</strong>al condition C(F, v, τ) = max[0, F − K]. σ F (t; T ) describes the volatility on<br />

the underly<strong>in</strong>g futures price whereas σ F v (t; T ) describes the “<strong>in</strong>stantaneous” rate of covariance<br />

between the futures price <strong>and</strong> the stochastic volatility process.<br />

This is basically the Heston (1993) model but with σF 2 (t; T ) <strong>and</strong> σ F v(t; T ) be<strong>in</strong>g functions<br />

of time <strong>in</strong>stead of constants (<strong>and</strong> the underly<strong>in</strong>g be<strong>in</strong>g a futures price <strong>in</strong>stead of a spot price).<br />

7


Hence, us<strong>in</strong>g the arguments <strong>and</strong> approach <strong>in</strong> Heston (1993), pp. 330-331 <strong>and</strong> his Appendix,<br />

one obta<strong>in</strong>s the follow<strong>in</strong>g option pric<strong>in</strong>g formula on the form of Black (1976),<br />

C(F, v, t) = e −r(τ−t) [F P 1 − KP 2 ] (17)<br />

where (see equation (18) <strong>in</strong> Heston (1993))<br />

P j = 1 2 + 1 ∫ [<br />

]<br />

∞<br />

e −iφ ln[K] f j (x, v, t; φ)<br />

Re<br />

dφ , j = 1, 2 (18)<br />

π<br />

iφ<br />

0<br />

<strong>and</strong> where x = ln F <strong>and</strong> f j (·; φ), j = 1, 2, denote characteristic functions that are obta<strong>in</strong>ed<br />

as solutions to PDE’s with term<strong>in</strong>al condition e iφx . S<strong>in</strong>ce coefficients <strong>in</strong> the particular PDE’s<br />

are aff<strong>in</strong>e <strong>and</strong> the term<strong>in</strong>al conditions are exponential-aff<strong>in</strong>e, the solutions for f j , j = 1, 2, are<br />

exponential-aff<strong>in</strong>e <strong>and</strong> on the form<br />

f j (x, v, t; φ) = e a j(t)+b j (t)v+iφx , j = 1, 2 (19)<br />

where a j (t) <strong>and</strong> b j (t) are solutions to the ord<strong>in</strong>ary differential equations<br />

b ′ j (t) = ( 1<br />

2 φ2 − u j iφ ) σ 2 F (t; T ) + (k j(t) − iφσ F v (t; T )) b j (t) − 1 2 σ2 v (b j (t)) 2 , b j (τ) = 0 (20)<br />

<strong>and</strong><br />

a ′ j(t) = −θb j (t) , a j (τ) = 0 (21)<br />

with u 1 = 1 2 , u 2 = − 1 2 , k 1(t) = κ − σ F v (t; T ), <strong>and</strong> k 2 (t) = κ.<br />

2.4 The estimation approach<br />

The estimation approach is based on formulat<strong>in</strong>g the model <strong>in</strong> state space form so that the<br />

model is represented by a measurement equation, <strong>and</strong> a transition equation. In our case<br />

the state space model is non-l<strong>in</strong>ear <strong>and</strong> the specific approach relies on a general state space<br />

estimation approach suggested orig<strong>in</strong>ally by Chen <strong>and</strong> Scott (1993) for estimation of models<br />

of the term structure of <strong>in</strong>terest rates.<br />

The data are <strong>in</strong> the follow<strong>in</strong>g observed at equidistant time po<strong>in</strong>ts, t n , n = 0, . . . , N, <strong>and</strong><br />

∆ = t n+1 − t n . The transition equation describes the dynamics of the basic unobserved<br />

state-variables. In our case, there are three state-variables which can be summarized by the<br />

three-dimensional state-vector X n = (p tn , δ tn , v tn ) where p t = log P t . The transition equation<br />

is given by the discrete-time solution to the stochastic differential system <strong>in</strong> (1), (2), <strong>and</strong> (3)<br />

between the observation time po<strong>in</strong>ts.<br />

8


The measurement equation relates the state-variables to the observed (log-) futures prices<br />

<strong>and</strong> option prices. Follow<strong>in</strong>g the approach <strong>in</strong> Chen <strong>and</strong> Scott (1993), we will assume that three<br />

prices on soybean cont<strong>in</strong>gent claims are observed without measurement error. In particular,<br />

this allows one to filter out the three unobserved state-variables by solv<strong>in</strong>g three equations<br />

with three unknowns at all observation time po<strong>in</strong>ts. Let the vector Y n denote the panel-data<br />

observation of (log-) futures prices <strong>and</strong> options prices at time t n , n = 0, . . . , N.<br />

We have<br />

Y ′ n = (Y ′<br />

(1),n , Y ′<br />

(2),n ) where Y (1),n is the three-dimensional vector of contract prices observed<br />

without measurement error while Y (2),n is the m-dimensional vector of contract prices that are<br />

observed with measurement error. Formally, the measurement equation is given by<br />

⎛ ⎞ ⎛ ⎞ ⎛ ⎞<br />

⎝ Y (1),n<br />

⎠ = ⎝ f (1),n(X n )<br />

⎠ + ⎝ 0 ⎠ (22)<br />

Y (2),n f (2),n (X n ) ɛ n<br />

where 0 is a three-dimensional vector of zeros <strong>and</strong> ɛ n , n = 0, 1, . . . , N, are serially uncorrelated<br />

measurement error terms that are assumed identical m-dimensional normally distributed with<br />

mean zero <strong>and</strong> variance-covariance matrix Γ; f (1),n <strong>and</strong> f (2),n are a three-dimensional valued<br />

function <strong>and</strong> an m-dimensional valued function, respectively, that return the relevant theoretical<br />

value of the logarithm to the futures prices <strong>and</strong> options prices, as implied by the futures<br />

<strong>and</strong> options pric<strong>in</strong>g expressions <strong>in</strong> (10) <strong>and</strong> (17).<br />

In our implementation, we assume that a short maturity at-the-money call option, the<br />

correspond<strong>in</strong>g futures price, <strong>and</strong> the futures price on a long maturity futures contract are<br />

observed without measurement error; i.e. these are the contract prices that enter Y (1),n at the<br />

different sample dates. (Y (2),n has dimension m = 10 <strong>in</strong> our implementation s<strong>in</strong>ce six (log-)<br />

futures prices <strong>and</strong> four option prices are assumed observed with measurement noise.) Hav<strong>in</strong>g<br />

observed a total of three futures <strong>and</strong> option prices without measurement noise, it is possible to<br />

filter out the three-dimensional state-vector X n by simply <strong>in</strong>vert<strong>in</strong>g the relationship between<br />

Y (1),n <strong>and</strong> X n , as given by the pric<strong>in</strong>g expressions <strong>in</strong> (10) <strong>and</strong> (17) that are the entries <strong>in</strong> f (1),n .<br />

Specifically, we have the relationship<br />

X n = f −1<br />

(1),n (Y (1),n). (23)<br />

Whenever we refer to the value of X n <strong>in</strong> the follow<strong>in</strong>g, it is assumed that X n is observed<br />

through the filter<strong>in</strong>g equation (23).<br />

Let ψ denote the set of parameters enter<strong>in</strong>g the description of state-variable dynamics <strong>in</strong><br />

(1), (2), <strong>and</strong> (3) <strong>and</strong> <strong>in</strong> the pric<strong>in</strong>g results <strong>in</strong> (10) <strong>and</strong> (17). We are <strong>in</strong>terested <strong>in</strong> estimat<strong>in</strong>g the<br />

9


parameters <strong>in</strong> ψ as well as the variance-covariance matrix Γ of the measurement error terms <strong>in</strong><br />

the measurement equation (22). Follow<strong>in</strong>g Chen <strong>and</strong> Scott (1993), the log-likelihood function<br />

has the follow<strong>in</strong>g form:<br />

l(Y 0 , Y 1 , . . . , Y N ; ψ, Γ) = ∑ [<br />

N<br />

∣<br />

n=1<br />

log p(X n |X n−1 ) − log<br />

∣<br />

∣ ∂f (1),n<br />

∂X<br />

− m 2 log(2π) − 1 2 log |Γ| − 1 2 ɛ′ nΓ −1 ɛ n<br />

]<br />

.<br />

The first term on the right-h<strong>and</strong> side of (24) is the log-likelihood value of the implied statevariables<br />

as filtered through the relationship <strong>in</strong> (23), <strong>and</strong> p(X n |X n−1 ), n = 1, . . . , N, denote<br />

the relevant conditional densities of the Markovian discrete-time solution to the stochastic<br />

differential equation system (1), (2), <strong>and</strong> (3) which describes the transition equation. 9<br />

second term on the right-h<strong>and</strong> side of (24) is the Jacobian determ<strong>in</strong>ant of the transformation<br />

between the observed contract prices <strong>in</strong> Y (1),n <strong>and</strong> X n . The last terms on the right-h<strong>and</strong> side<br />

of (24) are the likelihood of the normally distributed measurement errors.<br />

The estimation results obta<strong>in</strong>ed <strong>in</strong> the follow<strong>in</strong>g analysis is obta<strong>in</strong>ed by numerical maximization<br />

of the log-likelihood function <strong>in</strong> (24).<br />

(24)<br />

The<br />

However, s<strong>in</strong>ce the conditional densities<br />

p(X n |X n−1 ), n = 1, . . . , N, are not known <strong>in</strong> closed-form analytical form, we have chosen<br />

to rely on a quasi maximum likelihood approach where these conditional densities are substituted<br />

by densities from the three-dimensional normal distribution; the appropriate first- <strong>and</strong><br />

second order moments follow from Lemma 1, as stated <strong>and</strong> proved <strong>in</strong> the appendix. Likewise,<br />

<strong>in</strong> our calculation of st<strong>and</strong>ard errors on parameter estimates, we follow the general quasi maximum<br />

likelihood approach suggested orig<strong>in</strong>ally by White (1982); see also e.g. Hamilton (1994),<br />

pp. 126,145.<br />

The concrete implementation <strong>and</strong> optimization <strong>in</strong>volves many numerical considerations<br />

<strong>and</strong>, therefore, we have allocated the follow<strong>in</strong>g subsection to this issue.<br />

However, at this<br />

po<strong>in</strong>t it can be noted that the variance-covariance matrix of the measurement errors, Γ, can<br />

be concentrated out of the log-likelihood function.<br />

From the first order conditions of the<br />

maximization of (24), it is thus seen that the estimator of Γ (for a given set of parameters ψ)<br />

has the form<br />

ˆΓ = 1 N<br />

N∑<br />

ɛ n ɛ ′ n = 1 N<br />

n=1<br />

N∑ (<br />

Y(2),n − f (2),n (X n ) ) ( Y (2),n − f (2),n (X n ) ) ′<br />

n=1<br />

(25)<br />

9 Note that the <strong>in</strong>volved stochastic differential equation system is <strong>in</strong>homogeneous due to the seasonality<br />

feature of the model, though, the time-dependence of the relevant conditional densities is suppressed <strong>in</strong> (24).<br />

10


<strong>and</strong>, N ˆΓ ∼ W m (N, Γ) where W m (N, Γ) is the m-dimensional Wishart distribution with N degrees<br />

of freedom. Substitut<strong>in</strong>g the expression for Γ <strong>in</strong> (25) <strong>in</strong>to the log-likelihood function (24)<br />

simplifies the numerical optimization problem s<strong>in</strong>ce we only have to determ<strong>in</strong>e the maximum<br />

likelihood parameter values of ψ with a numerical optimization rout<strong>in</strong>e. The numerical optimization<br />

is done <strong>in</strong> GAUSS us<strong>in</strong>g the method of Broyden, Fletcher, Goldfarb <strong>and</strong> Shanno<br />

(BFGS). We use the st<strong>and</strong>ard GAUSS options <strong>in</strong> the numerical BFGS optimization algorithm.<br />

2.5 Numerical implementation<br />

In our implementation of the above estimation approach we observe data at a weekly frequency<br />

<strong>and</strong> at a total of 753 sample dates. At each sample date we have a panel-observation consist<strong>in</strong>g<br />

of eight (log-) futures prices <strong>and</strong> five option prices. The evaluation of the log-likelihood<br />

function, hence, <strong>in</strong>volves the evaluation of a large number of theoretical futures <strong>and</strong> options<br />

prices us<strong>in</strong>g the formulas stated <strong>in</strong> (10) <strong>and</strong> (17). In this section we will briefly describe how<br />

these pric<strong>in</strong>g expressions are implemented <strong>and</strong> how the log-likelihood function is calculated <strong>in</strong><br />

a relatively fast but still accurate way. Us<strong>in</strong>g the implementation described <strong>in</strong> the follow<strong>in</strong>g,<br />

the evaluation of a s<strong>in</strong>gle value of the log-likelihood function <strong>in</strong> GAUSS takes about one m<strong>in</strong>ute<br />

on a Pentium III 500 MHz computer while the evaluation of a s<strong>in</strong>gle option price on average<br />

is done <strong>in</strong> about one hundredth of a second. A s<strong>in</strong>gle iteration of the BFGS algorithm takes<br />

about one hour, <strong>and</strong> the parameter estimates presented <strong>in</strong> a subsequent section is obta<strong>in</strong>ed<br />

us<strong>in</strong>g a total of about three weeks of computer time.<br />

The evaluation of a futures price as <strong>in</strong> (10) requires solv<strong>in</strong>g the system of ord<strong>in</strong>ary differential<br />

equations <strong>in</strong> (12) <strong>and</strong> (13). The numerical solution to (12) <strong>and</strong> (13) is implemented us<strong>in</strong>g<br />

the classical Runge-Kutta method of fourth order accuracy comb<strong>in</strong>ed with Richardson extrapolation.<br />

The discrete time steps used <strong>in</strong> the classical Runge-Kutta method are always chosen<br />

so as not to exceed five days or, equivalently, (5/365.25=) 0.01369 years. In the Richardson<br />

extrapolation this solution is comb<strong>in</strong>ed with the Runge-Kutta solution with half as many time<br />

steps by us<strong>in</strong>g the relevant Richardson extrapolation formula for numerical methods of fourth<br />

order accuracy.<br />

The evaluation of an option price as <strong>in</strong> (17) requires the numerical evaluation of <strong>in</strong>tegrals of<br />

the form <strong>in</strong> (18). The numerical <strong>in</strong>tegration is implemented by the Gauss-Laguerre quadrature<br />

formula us<strong>in</strong>g twenty po<strong>in</strong>ts <strong>in</strong> the evaluation of the <strong>in</strong>tegr<strong>and</strong> function. Each evaluation of<br />

such an <strong>in</strong>tegral thus requires solv<strong>in</strong>g numerically the system of ord<strong>in</strong>ary differential equations<br />

11


<strong>in</strong> (20) <strong>and</strong> (21) for twenty different values of φ as well as simultaneously solv<strong>in</strong>g the ord<strong>in</strong>ary<br />

differential equation for B(·) <strong>in</strong> (12) s<strong>in</strong>ce the coefficients σ 2 F (·) <strong>and</strong> σ F v(·) <strong>in</strong> (20) are<br />

functions of B(·). This system of forty one (= 20 + 20 + 1) first order differential equations<br />

is solved simultaneously us<strong>in</strong>g the classical Runge-Kutta method comb<strong>in</strong>ed with Richardson<br />

extrapolation follow<strong>in</strong>g exactly the same procedure as described for futures prices above.<br />

Evaluation of the log-likelihood function <strong>in</strong> (24) requires <strong>in</strong>vert<strong>in</strong>g the relationship between<br />

the unobserved state-vector X n <strong>and</strong> the observed prices <strong>in</strong> Y (1),n , as described conceptually by<br />

the filter<strong>in</strong>g equation (23). In our implementation, we assume that a short maturity at-themoney<br />

call option, the correspond<strong>in</strong>g futures price, <strong>and</strong> the futures price on a long maturity<br />

futures contract are observed without measurement error.<br />

The <strong>in</strong>version is carried out by<br />

first solv<strong>in</strong>g for the implied value of the volatility, v tn , <strong>in</strong> the call option price by solv<strong>in</strong>g one<br />

equation with one unknown by the Newton-Raphson method (<strong>in</strong> this step we use that the<br />

underly<strong>in</strong>g futures price is observed without noise). In solv<strong>in</strong>g for the implied volatility by the<br />

Newton-Raphson method, we use that<br />

[<br />

∂C<br />

∂v = e−r(τ−t) F ∂P 1<br />

∂v − K ∂P ]<br />

2<br />

∂v<br />

<strong>and</strong><br />

∂P j<br />

∂v = 1 π<br />

∫ ∞<br />

0<br />

Re<br />

[<br />

b j (t) e−iφ ln[K] f j (x, v, t; φ)<br />

iφ<br />

]<br />

(26)<br />

dφ , j = 1, 2 (27)<br />

which follows by differentiat<strong>in</strong>g the call option expression <strong>in</strong> (17) <strong>and</strong> (18) for a fixed value<br />

of the underly<strong>in</strong>g futures price. Evaluat<strong>in</strong>g the relevant derivatives <strong>in</strong> this step thus require<br />

evaluation of <strong>in</strong>tegrals of the same form as <strong>in</strong> the evaluation of options prices, <strong>and</strong> the same<br />

procedure is applied <strong>in</strong> this case. 10 Hav<strong>in</strong>g solved for the implied value of the volatility statevariable,<br />

we use the exponential aff<strong>in</strong>e structure of the futures prices (i.e. the log-futures prices<br />

are aff<strong>in</strong>e functions of the state-variables) to solve for the values of the two rema<strong>in</strong><strong>in</strong>g statevariables<br />

p tn <strong>and</strong> δ tn by simply solv<strong>in</strong>g two l<strong>in</strong>ear equations with two unknowns. Furthermore,<br />

the Jacobian determ<strong>in</strong>ant of the transformation between Y (1),n <strong>and</strong> X n is calculated as a byproduct<br />

of the numerical calculations <strong>in</strong> this <strong>in</strong>version approach s<strong>in</strong>ce we have that<br />

∂f (1),n<br />

∣ ∂X ∣ = ∂C<br />

∂v (D(t n; T s ) − D(t n ; T l )) (28)<br />

10 It can be noted that the relevant system of forty one first order differential equations must only be solved<br />

once at each sample date <strong>in</strong> order to determ<strong>in</strong>e the relevant derivatives <strong>in</strong> the different Newton-Raphson steps<br />

as well as the calculation of all option prices with the same maturity.<br />

12


where T s <strong>and</strong> T l refer to the relevant maturities of the <strong>in</strong>volved short <strong>and</strong> long futures contract<br />

<strong>and</strong> D(·) is def<strong>in</strong>ed <strong>in</strong> (11).<br />

F<strong>in</strong>ally, the relevant conditional means <strong>and</strong> conditional variances that enter the normal<br />

density functions <strong>in</strong> the quasi maximum likelihood approach are determ<strong>in</strong>ed by solv<strong>in</strong>g numerically<br />

the <strong>in</strong>tegrals <strong>in</strong> (30) <strong>and</strong> (31), as stated <strong>in</strong> Lemma 1 <strong>in</strong> the appendix. S<strong>in</strong>ce data are<br />

observed at a weekly frequency we have ∆ = 1/52 = 0.01923 years. The numerical <strong>in</strong>tegration<br />

is <strong>in</strong> this case done by us<strong>in</strong>g the trapezoidal rule <strong>and</strong> us<strong>in</strong>g the second-order Runge-Kutta<br />

method (Heun’s method) with one time step to solve the differential equation <strong>in</strong> (32) <strong>and</strong><br />

obta<strong>in</strong> the value of Φ <strong>in</strong> the right-end-po<strong>in</strong>t of the <strong>in</strong>tegrals. Many of the entries <strong>in</strong> (30) <strong>and</strong><br />

(31) can be calculated <strong>in</strong> explicit analytical form, but not all, <strong>and</strong> the suggested approximation<br />

is very accurate compared to for example a left-end-po<strong>in</strong>t approximation of the <strong>in</strong>tegrals. The<br />

later approximation corresponds to the case where the relevant normal densities <strong>in</strong> the quasi<br />

maximum likelihood approach are obta<strong>in</strong>ed by an Euler approximation of the cont<strong>in</strong>uous-time<br />

differential equation system <strong>in</strong> (1), (2), <strong>and</strong> (3). In our estimations we orig<strong>in</strong>ally obta<strong>in</strong>ed<br />

start<strong>in</strong>g parameter values us<strong>in</strong>g such an Euler approximation, but with the specific data the<br />

estimates <strong>and</strong> the likelihood of observations are not ignorably altered by go<strong>in</strong>g to a more<br />

accurate approximation of the relevant moments <strong>in</strong> the conditional densities. 11<br />

3 Empirical results<br />

In this section we will describe the soybean futures <strong>and</strong> options data <strong>and</strong>, subsequently, present<br />

the formal estimation results based on the model <strong>and</strong> estimation approach presented above.<br />

3.1 Data<br />

The data consists of weekly observations of soybean futures prices <strong>and</strong> options on soybean<br />

futures from CBOT <strong>in</strong> the period from the beg<strong>in</strong>n<strong>in</strong>g of soybean options trad<strong>in</strong>g on CBOT <strong>in</strong><br />

October 1984 <strong>and</strong> to March 1999. The futures prices <strong>and</strong> option prices <strong>in</strong>volved are settlement<br />

prices at CBOT for Wednesdays (or Tuesday if Wednesday is unavailable). In fact, settlement<br />

prices are available on a daily basis but the weekly sampl<strong>in</strong>g frequency is chosen ma<strong>in</strong>ly to<br />

reduce problems due to microstructure issues such as daily price limits imposed by CBOT.<br />

11 In particular, the likelihood is significantly affected for values of the volatility state-variable be<strong>in</strong>g very close<br />

to zero.<br />

13


The Futures data<br />

The CBOT soybean futures have seven expiration months <strong>in</strong> the sampl<strong>in</strong>g period: January,<br />

March, May, July, August, September, <strong>and</strong> November. At any particular date, more than<br />

seven soybean futures contracts may be traded s<strong>in</strong>ce for example futures with expiration <strong>in</strong><br />

July this year <strong>and</strong> next year may be traded simultaneously. The maximum number of futures<br />

contracts that are open at all the dates <strong>in</strong> our sample is eight. In order to have same number<br />

of futures observations at each sample date, we have thus chosen to only <strong>in</strong>clude the eight<br />

futures contracts with the shortest time to expiration at any sample date.<br />

In the estimation approach we need the times to maturity for the different futures contracts<br />

at the different sample dates. The contractual terms at CBOT specify that physical delivery<br />

(<strong>in</strong> the form of an <strong>in</strong>ventory receipt) may occur at any bus<strong>in</strong>ess day throughout the expiration<br />

month <strong>and</strong>, hence, that the last delivery day is the last bus<strong>in</strong>ess day of the month. However,<br />

the last trad<strong>in</strong>g day of the futures contract is the seventh bus<strong>in</strong>ess day preced<strong>in</strong>g the last<br />

bus<strong>in</strong>ess day of the delivery month. S<strong>in</strong>ce the last trad<strong>in</strong>g day is the last day that a contract<br />

can <strong>in</strong> pr<strong>in</strong>ciple be closed at CBOT without physical delivery, we have chosen this day as the<br />

expiration date when construct<strong>in</strong>g the times to maturity for the <strong>in</strong>volved futures contracts <strong>in</strong><br />

the estimations presented <strong>in</strong> the next section.<br />

Table 1 provides summary statistics with respect to the <strong>in</strong>volved soybean futures contracts.<br />

[ INSERT TABLE 1 ABOUT HERE ]<br />

The table states the average number of days to maturity <strong>and</strong> mean <strong>and</strong> st<strong>and</strong>ard deviation<br />

of all the soybean futures prices <strong>in</strong> the dataset. In addition, the tables provide summary<br />

statistics for the futures contracts categorized <strong>in</strong>to expirations months as well as the time to<br />

maturity of the contracts. In the follow<strong>in</strong>g, the term<strong>in</strong>ology “1. Maturity” is used as notation<br />

for the futures contract that has the shortest time to maturity at a given sample date; the “2.<br />

Maturity” represents the futures contract with the second shortest time to maturity, <strong>and</strong> so<br />

on.<br />

The table suggests at least three features of the soybean futures prices that are captured by<br />

the model <strong>in</strong> section 2. (i) Futures prices display a seasonal pattern s<strong>in</strong>ce the soybean futures<br />

price is on average high <strong>in</strong> July prior to the US harvest<strong>in</strong>g <strong>and</strong> low <strong>in</strong> November after the US<br />

harvest<strong>in</strong>g. (ii) The variability of futures prices also seem to have a seasonal pattern s<strong>in</strong>ce the<br />

st<strong>and</strong>ard deviations for the futures prices, when grouped <strong>in</strong>to expiration months, display the<br />

same time pattern as for the level of futures prices. And, (iii) the variability of long futures<br />

14


prices seems lower than that of short futures prices. The lower variation of futures prices with<br />

a long time to maturity compared to futures prices with a short time to maturity is suggested<br />

by the different st<strong>and</strong>ard deviations for the futures prices grouped <strong>in</strong>to their time to maturity<br />

<strong>in</strong> Table 1. The tabulated st<strong>and</strong>ard deviations are thus monotonically decreas<strong>in</strong>g for <strong>in</strong>creas<strong>in</strong>g<br />

maturities. This pattern is consistent with the so-called “Samuelson hypothesis” that predicts<br />

that long futures prices are less volatile than futures prices on contracts with short time to<br />

maturity.<br />

The time series aspects of the soybean futures data are illustrated <strong>in</strong> Figure 1.<br />

[ INSERT FIGURE 1 ABOUT HERE ]<br />

In the figure we have graphed the time-series for the futures contracts with the shortest time to<br />

maturity <strong>and</strong> the futures contracts with the longest time to maturity over the sample period. A<br />

visual <strong>in</strong>spection of the figure suggests that the time series of the futures price on the shortest<br />

maturity contract is more volatile than the futures price on the longest maturity contract. In<br />

particular, if one enter a horizontal l<strong>in</strong>e at the overall futures price mean at 627.49 cents per<br />

bushel (cf. Table 1) <strong>in</strong> the figure, the deviations from this overall mean is more significant<br />

for the short maturity futures price over the sample period. Aga<strong>in</strong>, this is <strong>in</strong> l<strong>in</strong>e with the<br />

“Samuelson hypothesis” <strong>and</strong> a feature of the soybean futures prices which is captured <strong>in</strong> the<br />

formal model<strong>in</strong>g <strong>in</strong> section 2 by <strong>in</strong>clud<strong>in</strong>g a mean-revert<strong>in</strong>g convenience yield.<br />

The options data<br />

The <strong>in</strong>volved CBOT soybean option contracts are options written on soybean futures prices.<br />

The relevant futures contracts expire <strong>in</strong>: January, March, May, July, August, September,<br />

<strong>and</strong> November. The option contract expires around one month prior to the expiration of the<br />

underly<strong>in</strong>g futures contract. Specifically, accord<strong>in</strong>g to the CBOT contract specification, the<br />

last trad<strong>in</strong>g day is the Friday which precedes by at least five bus<strong>in</strong>ess days the last bus<strong>in</strong>ess<br />

day of the month preced<strong>in</strong>g the option month. The expiration date is the Saturday follow<strong>in</strong>g<br />

the last trad<strong>in</strong>g day.<br />

Both call options <strong>and</strong> put options are traded. Also, at any particular sample date a variety<br />

of options contracts, which differ with respect to the underly<strong>in</strong>g futures <strong>and</strong> with respect to<br />

the strike prices, are quoted. However, <strong>in</strong> our sample we have only <strong>in</strong>cluded five call option<br />

contracts at each sample date. The specific call option contracts are selected <strong>in</strong> order to<br />

represent very liquid contracts. The sample thus <strong>in</strong>cludes three call option contracts that<br />

15


have the shortest possible time to maturity (though, we ignore very short contracts so that<br />

all the considered options have at least 15 days to maturity). The three call options with a<br />

short maturity are an at-the-money (ATM) option, an out-of-the-money (OTM) option, <strong>and</strong><br />

an <strong>in</strong>-the-money (ITM) option. The ATM option is chosen as the call option that has a strike<br />

price as close as possible to the current underly<strong>in</strong>g futures price. The OTM call option is the<br />

quoted call option with a strike price just higher than the ATM call option <strong>and</strong>, likewise, the<br />

ITM call option is the quoted call option with a strike price just lower than the ATM call<br />

option. Furthermore, our sample <strong>in</strong>cludes two longer term ATM options that are written on<br />

the next matur<strong>in</strong>g futures prices.<br />

The CBOT options on soybean futures are American-style. S<strong>in</strong>ce our pric<strong>in</strong>g formulas are<br />

relevant only for European-style options, the option data are adjusted for the early exercise<br />

premium. The early exercise premia is estimated us<strong>in</strong>g a st<strong>and</strong>ard Black-Scholes framework<br />

<strong>and</strong> apply<strong>in</strong>g a tr<strong>in</strong>omial lattice model comb<strong>in</strong>ed with Richardson extrapolation, follow<strong>in</strong>g<br />

Broadie <strong>and</strong> Detemple (1996). 12 The relevant <strong>in</strong>terest rates are obta<strong>in</strong>ed by <strong>in</strong>terpolation of<br />

3 months, 6 months, <strong>and</strong> 1 year to maturity T-bills available from the Federal Reserve H.15<br />

release. The rates are quoted on a banker’s discount basis <strong>and</strong> converted appropriately; see<br />

Jarrow (1996), chapter 2. The American-style CBOT option prices m<strong>in</strong>us the estimated early<br />

exercise premia are the European-style <strong>in</strong>put data used <strong>in</strong> the empirical analysis.<br />

Table 2 provides summary statistics on the <strong>in</strong>volved <strong>in</strong>put call option prices.<br />

[ INSERT TABLE 2 ABOUT HERE ]<br />

Instead of exhibit<strong>in</strong>g summary statistics of the raw option prices, we have chosen to represent<br />

the option data by implied volatilities <strong>in</strong> Table 2. The relevant implied volatilities of the<br />

<strong>in</strong>volved options on soybean futures are backed out us<strong>in</strong>g the Black (1976) formula. Three<br />

features of the option data can be po<strong>in</strong>ted out: (i) The implied volatilities have a clear seasonal<br />

pattern, (ii) the implied volatilities for the ATM options are slightly <strong>in</strong>creas<strong>in</strong>g as a function<br />

of time to maturity, <strong>and</strong> (iii) the volatilities on the short maturity contract suggest a “smile”<br />

or “smirk” pattern. Note, however, that the “smirk” is downward slop<strong>in</strong>g as a function of the<br />

strike price <strong>in</strong> contrast to the upward slop<strong>in</strong>g “smirk” <strong>in</strong> equity <strong>in</strong>dex options observed after<br />

the October 1987 market crash; see, e.g., Rub<strong>in</strong>ste<strong>in</strong> (1994).<br />

In Figure 2 the time series properties of the implied volatilities for the shortest ATM option<br />

are displayed.<br />

12 The results <strong>in</strong> Broadie <strong>and</strong> Detemple (1996) suggest that this method is efficient with respect to accuracy.<br />

16


[ INSERT FIGURE 2 ABOUT HERE ]<br />

As suggested by Figure 2, the implied volatilities are quite variable over time. This, comb<strong>in</strong>ed<br />

with the clear seasonal pattern <strong>in</strong> implied volatilities displayed <strong>in</strong> Table 2, is our basic<br />

motivation for the formal modell<strong>in</strong>g of the stochastic volatility <strong>in</strong> section 2.<br />

3.2 Estimation results<br />

In this section we will present <strong>and</strong> discuss the estimation results based on the above data set<br />

<strong>and</strong> the estimation approach described <strong>in</strong> section 2.4. We start by discuss<strong>in</strong>g the full-sample<br />

estimates but, subsequently, we also present estimation results where the data is split <strong>in</strong>to two<br />

sub-samples <strong>in</strong> order to <strong>in</strong>vestigate the parameter stability of the model over time.<br />

The parameter estimates obta<strong>in</strong>ed from the state space estimation approach <strong>and</strong> the full<br />

sample of observations are tabulated <strong>in</strong> Table 3.<br />

[ INSERT TABLE 3 ABOUT HERE ]<br />

In the follow<strong>in</strong>g discussion of the full-sample parameter estimates <strong>in</strong> Table 3, we will start<br />

by focus<strong>in</strong>g on the structural parameters of the soybean spot price volatility process <strong>and</strong> the<br />

convenience yield process parameters <strong>and</strong>, subsequently, we discuss the implications of the<br />

estimated measurement noise structure.<br />

<strong>Stochastic</strong> volatility <strong>and</strong> seasonality<br />

The volatility of the implied soybean spot price is determ<strong>in</strong>ed by the process v t <strong>and</strong> the seasonal<br />

function ν(t). In particular, as <strong>in</strong>ferred from the spot price dynamics <strong>in</strong> (1), or the risk-neutral<br />

analog <strong>in</strong> (4), e 2ν(t) v t is the rate of variance on the soybean spot price <strong>and</strong>, hence, e ν(t)√ v t is<br />

the spot price volatility. S<strong>in</strong>ce the seasonal function ν(t) captures any regular <strong>and</strong> systematic<br />

seasonal pattern <strong>in</strong> the spot price volatility, √ v t can be <strong>in</strong>terpreted as the seasonally adjusted<br />

spot price volatility. Under the risk-neutral measure, the long-run level for v t is estimated to be<br />

0.0679 (= θ/κ) which implies a value of the long-run seasonal adjusted volatility of about 26.1%<br />

(= √ 0.0679). Furthermore, the volatility process exhibits a strong degree of mean-reversion<br />

s<strong>in</strong>ce the parameter estimate of κ is significantly higher than zero. 13 The parameter estimate<br />

13 Throughout the discussion we will refer to an estimate be<strong>in</strong>g significantly differently from a given value if<br />

the given value is not with<strong>in</strong> the estimate plus/m<strong>in</strong>us two st<strong>and</strong>ard deviations (the usually applied asymptotic<br />

95% confidence <strong>in</strong>terval). The 95% confidence <strong>in</strong>terval for κ is thus [1.7400, 2.8016].<br />

17


of κ (= 2.2708) implies that the so-called half-life of shocks to the seasonal adjusted variance<br />

rate v t is 0.305 years (= log 2/κ), under the risk-neutral probability measure. Basically, this<br />

means that it takes about three to four months before a given shock <strong>in</strong> volatility (e.g. due to<br />

changes <strong>in</strong> harvest expectations or temporary changes <strong>in</strong> dem<strong>and</strong> for soybeans) is expected to<br />

have levelled off to half of its immediate effect. Under the true probability measure, however,<br />

the degree of mean-reversion is even stronger <strong>and</strong> estimated to be 5.0754 (= κ − λ v σ v ) which<br />

corresponds to a half-life of shocks of 0.137 years, or about one <strong>and</strong> a half months. Moreover,<br />

under the true probability measure the long-run level of the seasonal adjusted volatility, as<br />

reflected <strong>in</strong> the estimates <strong>in</strong> Table 3, is 17.4% (= √ θ/5.0754). These long-run volatility levels<br />

seem well <strong>in</strong> accordance with the overall average Black-Scholes implied volatility of 19.99%,<br />

as tabulated <strong>in</strong> the summary statistics <strong>in</strong> Table 2.<br />

The seasonal variation <strong>in</strong> volatilities is captured by the seasonal function ν(t) def<strong>in</strong>ed <strong>in</strong> (8).<br />

This function is a sum of self-repeat<strong>in</strong>g trigonometric functional terms. In our estimation we<br />

have on beforeh<strong>and</strong> limited the number of terms to two <strong>in</strong> order to keep the model sufficiently<br />

parsimonious for numerical optimization to be possible <strong>and</strong>, hence, <strong>in</strong> our implementation this<br />

function is described by a total of four parameters: ν 1 , ν1 ∗, ν 2, <strong>and</strong> ν2 ∗.14<br />

All the seasonal<br />

parameters are estimated to be significantly different from zero. The seasonal pattern <strong>in</strong><br />

volatilities that the parameter estimates reflect is illustrated <strong>in</strong> Figure 3 where the seasonal<br />

function ν(t) is plotted.<br />

[ INSERT FIGURE 3 ABOUT HERE ]<br />

The figure illustrates that the seasonal function has two local maxima <strong>and</strong> two local m<strong>in</strong>ima<br />

(though, this feature is much more clear for the convenience yield seasonal function α(t), as<br />

to be discussed below). The global maximum is achieved <strong>in</strong> late July about two months prior<br />

to the beg<strong>in</strong>n<strong>in</strong>g of the US harvest (which occur from mid to late September <strong>and</strong> through<br />

October). 15 The flower<strong>in</strong>g <strong>and</strong> pod fill<strong>in</strong>g of the plant, which is of crucial importance for<br />

the f<strong>in</strong>al seed yield <strong>and</strong> soybean production occur around this time, <strong>and</strong> this is thus a period<br />

14 We have estimated the model <strong>in</strong>clud<strong>in</strong>g additional terms <strong>in</strong> the seasonal functions ν(t) <strong>and</strong> α(t) (the analog<br />

for the convenience yield process) but keep<strong>in</strong>g all other parameters fixed at the estimated values <strong>in</strong> Table 3. This<br />

did not affect the estimated seasonal patterns substantially. Also, it may be noted that <strong>in</strong> a similar analysis with<br />

futures price observations <strong>and</strong> seasonality <strong>in</strong> convenience yields, Sørensen (2001) limits the relevant number of<br />

terms <strong>in</strong>volved <strong>in</strong> the seasonal function to two based on a formal criteria, the Akaike Information Criteria.<br />

15 The global maximum of the seasonal function ν(t) is reached on July 21 where the function value is 0.365<br />

while the global m<strong>in</strong>imum is reached on March 23 where the function value is -0.249.<br />

18


where adverse weather conditions may significantly affect next year’s US supply of soybeans.<br />

At this time of the year the volatility <strong>in</strong> soybeans is estimated to be 44% (= e 0.365 − 1)<br />

above “normal,” where normal refers to the case when ν(t) = 0. On the other h<strong>and</strong>, the<br />

soybean volatility is low <strong>in</strong> March prior to the US plant<strong>in</strong>g (which occur <strong>in</strong> May <strong>and</strong> June)<br />

<strong>and</strong> at its lowest the volatility is 22% (= 1 − e −0.249 ) below normal. As discussed above,<br />

the seasonal adjusted volatility reverts over time to a long-run level of around 17.4%. The<br />

estimated seasonal pattern, however, suggests that the mean-revert<strong>in</strong>g level is closer to 25.1%<br />

(= 1.44 × 17.4%) <strong>in</strong> July <strong>and</strong> 13.6% (= 0.78 × 17.4%) <strong>in</strong> March. These mean-revert<strong>in</strong>g levels<br />

are of the same magnitude, though slightly below, the average volatilities tabulated <strong>in</strong> Table 2.<br />

Note <strong>in</strong> addition that <strong>in</strong> Table 2 the m<strong>in</strong>imum average implied volatility is obta<strong>in</strong>ed for the<br />

options written on the May futures while the maximum average implied volatility is obta<strong>in</strong>ed<br />

for the option written on the September futures. But s<strong>in</strong>ce these option contracts expire one<br />

month prior <strong>in</strong> April <strong>and</strong> August, respectively, <strong>and</strong> s<strong>in</strong>ce the average times to maturity are<br />

78.49 days <strong>and</strong> 59.04 days, these option contracts are precisely the contracts that cover the<br />

relevant periods through late March <strong>and</strong> late July, respectively.<br />

Convenience yields <strong>and</strong> theory of storage<br />

The parameters estimated for the convenience yield process <strong>in</strong> (2), (5), <strong>and</strong> (7) also reveal a<br />

clear seasonal pattern. In particular, the convenience yield process reverts to a level that depends<br />

on the season, as reflected <strong>in</strong> the seasonal function α(t). The degree of tendency towards<br />

this “normal” seasonal level is described by the parameter β. Note that while the convenience<br />

yield drift term has this simple <strong>in</strong>terpretation under the risk neutral measure <strong>in</strong> (5), the dynamics<br />

are more complicated under the true probability measure s<strong>in</strong>ce the convenience yield<br />

risk premium is assumed proportional to the time-vary<strong>in</strong>g volatility (<strong>in</strong>clud<strong>in</strong>g the seasonality<br />

feature of volatility). However, s<strong>in</strong>ce the convenience yield risk premium parameter λ δ is estimated<br />

close to (<strong>and</strong> not significantly different from) zero, we will cont<strong>in</strong>ue the discussion as if<br />

λ δ = 0 or, equivalently, as if the risk-neutral dynamics <strong>and</strong> the true dynamics co<strong>in</strong>cide. The<br />

estimate of β is 0.8145 which imply a half-life of unexpected shocks to the convenience yield<br />

process of about 0.851 years (= log 2/β). Hence, the mean-revert<strong>in</strong>g tendency is seem<strong>in</strong>gly<br />

less strong than the similar feature <strong>in</strong> the volatility process, as discussed above. The seasonal<br />

function α(t) is described by a total of five parameters: α 0 , α 1 , α1 ∗, α 2, <strong>and</strong> α2 ∗ . These are all<br />

estimated to be significantly different from zero. The parameter α 0 is a constant <strong>and</strong> without<br />

19


a seasonality feature the long-run level for the convenience yield process would thus be 0.0751<br />

(= α 0 /β); we <strong>in</strong>terpret this as the “normal” level of the convenience yield when adjusted for<br />

seasonality.<br />

The seasonal pattern of the mean-revert<strong>in</strong>g level of the convenience yield is also illustrated<br />

<strong>in</strong> Figure 3. The seasonal variation <strong>in</strong> α(t) is very similar to the seasonal variation <strong>in</strong> ν(t),<br />

though amplitudes are different. However, it is much more clear that α(t) has two local maxima<br />

<strong>and</strong> two local m<strong>in</strong>ima. 16 Aga<strong>in</strong>, the global maximum is achieved <strong>in</strong> July about two months<br />

prior to the beg<strong>in</strong>n<strong>in</strong>g of the US harvest<strong>in</strong>g while the global m<strong>in</strong>imum is achieved <strong>and</strong> the end<br />

of the US harvest<strong>in</strong>g. While the US soybean production accounts for 50-60% of world soybean<br />

production other key producers are Argent<strong>in</strong>a <strong>and</strong>, especially, Brazil which are both located<br />

<strong>in</strong> the Southern Hemisphere. These countries account together for 20-30% of world soybean<br />

production. The Brazilian soybean harvest<strong>in</strong>g takes place <strong>in</strong> March <strong>and</strong> April <strong>and</strong> the local<br />

maximum <strong>in</strong> Figure 3 is reached about two months earlier, while the local m<strong>in</strong>imum is reached<br />

dur<strong>in</strong>g the end<strong>in</strong>g of the Brazilian soybean harvest<strong>in</strong>g. The seasonal pattern <strong>in</strong> convenience<br />

yields is thus related to the supply of soybeans. When supplies are low convenience yields are<br />

high, <strong>and</strong> vice versa. This is consistent with the theory of storage which predicts a negative<br />

relationship between <strong>in</strong>ventories <strong>and</strong> convenience yields. The basic idea is that the hold<strong>in</strong>g of<br />

physical <strong>in</strong>ventory of soybeans gives rise to a convenience yield from be<strong>in</strong>g able to profit from<br />

temporary soybean supply shortages or keep a production process runn<strong>in</strong>g especially when<br />

<strong>in</strong>ventories are low. Furthermore, the parameter estimates suggest that the soybean spot price<br />

<strong>and</strong> the convenience yield are positively correlated with a correlation coefficient of 0.3979.<br />

Follow<strong>in</strong>g the above k<strong>in</strong>d of reason<strong>in</strong>g, this is also consistent with the observation that when<br />

supplies <strong>and</strong> <strong>in</strong>ventories are scarce the equilibrium soybean spot price <strong>and</strong> the convenience<br />

yield are simultaneously high; <strong>and</strong> vice versa when supplies <strong>and</strong> <strong>in</strong>ventories are plenty.<br />

On the other h<strong>and</strong>, supporters of the theory of storage often view the convenience yield as<br />

an option to profit from temporary shortages of the particular commodity. This would suggest<br />

a positive correlation between the convenience yield <strong>and</strong> the volatility; however, this view is<br />

not supported by the parameter estimate of ρ 23 which is not significantly different from zero<br />

<strong>and</strong> even estimated negative.<br />

16 The global maximum of the seasonal function α(t) is reached on July 15 where the function value is 0.746,<br />

while the global m<strong>in</strong>imum is reached on October 30 where the function value is -0.723. A local maximum is<br />

reached on February 7, <strong>and</strong> a local m<strong>in</strong>imum is reached on April 16.<br />

20


Inverse leverage effects<br />

The correlation coefficient between the soybean spot price <strong>and</strong> the soybean volatility is estimated<br />

to be 0.4078. This reflects that the volatility tend to be high when soybean spot prices<br />

are high <strong>and</strong>, <strong>in</strong> which case, supplies/<strong>in</strong>ventories tend to be scarce. While this seems only<br />

natural <strong>and</strong> an <strong>in</strong>tuitively appeal<strong>in</strong>g feature for commodities like soybeans, this is different<br />

from equity markets where spot prices <strong>and</strong> volatilities are usually estimated to be negatively<br />

correlated; <strong>in</strong> equity markets this phenomenon is often referred to as the leverage effect. This<br />

is consistent with the observation that the volatility “smirk” <strong>in</strong> soybean option prices has the<br />

opposite slope of that implied from equity options after October 1987, as described <strong>in</strong> our data<br />

description. Note also that the estimated positive correlation between the spot price <strong>and</strong> the<br />

volatility seems <strong>in</strong>tuitively to support the observation that OTM call option should trade at<br />

higher implied volatilities. In particular, <strong>in</strong> order to have a positive payoff <strong>and</strong> value effect on<br />

the call option, the underly<strong>in</strong>g spot price must <strong>in</strong>crease <strong>and</strong>, s<strong>in</strong>ce the volatility <strong>in</strong> this case<br />

also tends to <strong>in</strong>crease, this will magnify the upward potential of the particular option (while<br />

keep<strong>in</strong>g downside risk limited).<br />

Risk premia <strong>and</strong> <strong>in</strong>vestment returns<br />

As seen from Table 3, <strong>in</strong> all cases the risk premia parameters are estimated with large st<strong>and</strong>ard<br />

errors. Only the risk premium on volatility risk is significantly different from zero, <strong>and</strong> the<br />

implications with respect to the long-run level for the seasonal adjusted volatility under the<br />

risk-neutral measure, as reflected <strong>in</strong> market prices, <strong>and</strong> the true probability measure was<br />

discussed earlier on.<br />

While no strong <strong>in</strong>ference can be drawn from the other risk premia<br />

estimates, it seems relevant to briefly consider the implications of the parameter po<strong>in</strong>t estimates<br />

for the implied excess returns on e.g. soybean futures positions. We will consider an <strong>in</strong>vestor<br />

who exposes himself to soybean price risk by go<strong>in</strong>g long <strong>in</strong> futures contracts <strong>and</strong> at the same<br />

time places an amount equal to the futures price on a safe bank account.<br />

S<strong>in</strong>ce the drift<br />

rate of the futures price is zero under the risk-neutral probability measure, the excess return<br />

on this <strong>in</strong>vestment policy is described by the drift rate of the futures price under the true<br />

probability measure. Us<strong>in</strong>g Ito’s lemma on the futures price expression <strong>in</strong> equation (10), <strong>and</strong><br />

by comparison with the risk-neutral dynamics of the futures price <strong>in</strong> (14), it is thus seen that<br />

the expected excess return on this <strong>in</strong>vestment position is<br />

[ ] dFt (T )<br />

(<br />

)<br />

E t = λ P e ν(t) + λ δ σ δ e ν(t) D(t; T ) + λ v σ v B(t; T ) v t (29)<br />

F t (T )<br />

21


where D(t; T ) <strong>and</strong> B(t, T ) are described <strong>in</strong> (11) <strong>and</strong> (12). Assume that we are look<strong>in</strong>g at the<br />

futures position at May 17 <strong>in</strong> a given year; this date is chosen as a po<strong>in</strong>t <strong>in</strong> time t where<br />

the seasonal component <strong>in</strong> volatility is zero (i.e. ν(t) = 0). Also, assume that the spot price<br />

volatility is √ v t = 0.20 (i.e. v t = 0.04). If the <strong>in</strong>volved futures contract is very short (i.e. t = T ),<br />

we have that D(t; t) = B(t; t) = 0 <strong>and</strong> that the position is only exposed to spot commodity<br />

price risk. Insert<strong>in</strong>g the parameter po<strong>in</strong>t estimates <strong>in</strong> Table 3 <strong>in</strong>to (29), the expected excess<br />

return <strong>in</strong> this case is negative <strong>and</strong> equal to -2.51%. Longer term futures prices are negatively<br />

related to the convenience yield (see equations (10) <strong>and</strong> (11)). Hence, s<strong>in</strong>ce the risk premium<br />

on convenience yield risk is estimated negative, longer term futures contracts will have a higher,<br />

<strong>and</strong> possibly positive, overall risk premium when us<strong>in</strong>g the po<strong>in</strong>t estimates <strong>in</strong> Table 3. Also,<br />

as seen by numerical <strong>in</strong>spection, futures prices at the po<strong>in</strong>t estimates are negatively related to<br />

volatility. S<strong>in</strong>ce the risk premium on volatility risk is estimated negative, this also implies that<br />

longer term futures contracts will have a higher, <strong>and</strong> possibly positive, overall risk premium.<br />

However, when solv<strong>in</strong>g (11) <strong>and</strong> (12) for D(t; T ) <strong>and</strong> B(t, T ), <strong>and</strong> by <strong>in</strong>sertion <strong>in</strong> (29), we f<strong>in</strong>d<br />

that the overall expected excess return is -1.95% if the <strong>in</strong>volved futures contract has time to<br />

maturity equal to half a year. Likewise, the overall expected excess return is -1.77% if the<br />

<strong>in</strong>volved futures contract has time to maturity equal to one year. 17 In all the cases considered<br />

above, the risk premium on the long futures <strong>in</strong>vestment position is negative. Furthermore, s<strong>in</strong>ce<br />

the futures price is expected to drift downwards this is also a case where the futures price is<br />

above the expected future spot price, <strong>and</strong> thus a situation usually referred to as “contango.” 18<br />

On the other h<strong>and</strong>, the only risk premia parameter which is significantly different from zero<br />

is λ v <strong>and</strong> if we set λ P = λ δ = 0, normal backwardation would be the situation. It is thus<br />

concluded that no strong <strong>in</strong>ference can be drawn from the risk premia po<strong>in</strong>t estimates with<br />

respect to the quantitative implications as well as the qualitative implications for expected<br />

soybean <strong>in</strong>vestment returns.<br />

Structure of measurement error covariance matrix<br />

The measurement noise term added <strong>in</strong> the state space specification of the model is estimated<br />

to have a covariance matrix as implied by the stated correlation matrix of measurement errors<br />

<strong>in</strong> Table 3. Thus, us<strong>in</strong>g the one-to-one relationship between the tabulated correlation matrix<br />

17 By solv<strong>in</strong>g (11) <strong>and</strong> (12) numerically, we <strong>in</strong> the particular cases f<strong>in</strong>d that B(t; t + 0.50) = -0.03151, D(t; t +<br />

0.50) = −0.41071, B(t; t + 1.00) = -0.03475, <strong>and</strong> D(t; t + 1.00) = -0.68403.<br />

18 see e.g. Hull (2000, p. 74) for this def<strong>in</strong>ition of contango versus normal backwardation.<br />

22


(with the relevant estimated st<strong>and</strong>ard deviation <strong>in</strong> the diagonal), it is straightforward to<br />

obta<strong>in</strong> the relevant covariance matrix, <strong>and</strong> vice versa. Moreover, applicable st<strong>and</strong>ard errors<br />

on the covariance matrix can be obta<strong>in</strong>ed by us<strong>in</strong>g the result stated <strong>in</strong> section 2.4 that N ˆΓ<br />

is Wishart distributed. In particular, hav<strong>in</strong>g N=753 observation dates the st<strong>and</strong>ard errors on<br />

the st<strong>and</strong>ard deviation estimates <strong>in</strong> the diagonal can be obta<strong>in</strong>ed as 0.0364 (= 1/ √ 753) times<br />

the true st<strong>and</strong>ard deviation parameter (which, as usual, may by approximated by the po<strong>in</strong>t<br />

estimate). S<strong>in</strong>ce we are only <strong>in</strong>terested <strong>in</strong> the overall structure of the measurement errors, we<br />

have only tabulated the po<strong>in</strong>t estimates on the covariance matrix <strong>in</strong> Table 3, <strong>and</strong>, <strong>in</strong> addition,<br />

we have chosen to represent the covariance matrix <strong>in</strong> the form of the implied correlation matrix<br />

<strong>in</strong> order to focus on a normalized measure of how the model errors co-vary which potentially<br />

could give some <strong>in</strong>dications <strong>and</strong> directions for how to improve the formal underly<strong>in</strong>g futures<br />

<strong>and</strong> option pric<strong>in</strong>g model.<br />

The measurement noise error vector has dimension ten, as described <strong>in</strong> the data description.<br />

The first four entries <strong>in</strong> the vector are related to option prices <strong>in</strong> the follow<strong>in</strong>g order: a short<br />

maturity <strong>in</strong>-the-money call option, a short maturity out-of-the-money call option, a medium<br />

maturity at-the-money call option, <strong>and</strong> a long maturity at-the-money call option. In general,<br />

the longer the maturity the less precise is the formal model <strong>in</strong> provid<strong>in</strong>g at-the-mark option<br />

prices, as can be <strong>in</strong>ferred from the higher st<strong>and</strong>ard deviations on the measurement error terms<br />

for long maturity contracts. Also, it is evident that the measurement errors on the medium<br />

maturity <strong>and</strong> long maturity call options are positively correlated so that if the medium maturity<br />

call option is underpriced by the theoretical option pric<strong>in</strong>g model then the long maturity call<br />

option contract will also tend to be underpriced; the po<strong>in</strong>t estimate of the relevant correlation<br />

coefficient is 0.7310. It is likewise seen that the measurement errors on the short <strong>in</strong>-the-money<br />

call option <strong>and</strong> out-of-the-money call option are negatively correlated. This could be consistent<br />

with, <strong>and</strong> thus <strong>in</strong>dicative of, an even steeper empirical volatility surface than produced by the<br />

formal option pric<strong>in</strong>g model with parameters as <strong>in</strong> Table 3 s<strong>in</strong>ce a relatively flat model volatility<br />

curve would tend to produce undervalued <strong>in</strong>-the-money options whenever out-of-the money<br />

options are overvalued, <strong>and</strong> vice versa.<br />

The last six entries <strong>in</strong> the measurement noise errors correspond to relative errors on futures<br />

prices with different times to maturity. Number five entry relates to the shortest time to<br />

maturity futures contract observed with measurement error, number six relates to the second<br />

shortest time to maturity futures contract observed with measurement error, <strong>and</strong> so on. In<br />

the estimation approach we have assumed that futures prices on a short maturity contract<br />

23


<strong>and</strong> the longest available futures contract are observed without measurement error. Hence, it<br />

is not surpris<strong>in</strong>g that short <strong>and</strong> long contracts are seem<strong>in</strong>gly fitted better by the model than<br />

medium maturity contracts, as reflected <strong>in</strong> the higher st<strong>and</strong>ard deviations on the imposed<br />

measurement errors for medium maturity contracts. Note that the st<strong>and</strong>ard deviations on the<br />

<strong>in</strong>dividual error terms on futures prices are estimated of the same numerical magnitude as <strong>in</strong><br />

Schwartz (1997), Schwartz <strong>and</strong> Smith (2000), <strong>and</strong> Sørensen (2001) who use only commodity<br />

futures data <strong>and</strong> Kalman filter<strong>in</strong>g estimation approaches. For example, Sørensen (2001) provides<br />

results for soybean futures where the measurement errors are assumed uncorrelated <strong>and</strong><br />

with identical st<strong>and</strong>ard deviation which he estimates to be 0.187. As apparent from the correlation<br />

matrix <strong>in</strong> Table 3, especially measurement errors for contracts that have maturities close<br />

to each other are highly positively correlated while the correlations decrease systematically for<br />

contracts hav<strong>in</strong>g <strong>in</strong>creas<strong>in</strong>gly dispersed maturities.<br />

F<strong>in</strong>ally, the estimated correlations <strong>in</strong>dicate slight co-variation between option prices <strong>and</strong><br />

futures prices measurement errors. In particular, the measurement errors related to short maturity<br />

options have higher correlations with the measurement errors related to short maturity<br />

futures while the medium/long maturity option measurement errors are more correlated with<br />

measurement errors on futures prices of similar maturities. The correlations are <strong>in</strong> general<br />

positive, however, no clear pattern is evident s<strong>in</strong>ce the measurement error related to short maturity<br />

<strong>in</strong>-the-money call options is negatively correlated with the measurement error on futures<br />

prices of similar maturity while the short maturity out-of-the-money call option measurement<br />

error at the same time is positively correlated with the same futures price measurement error.<br />

Structural changes<br />

In order to provide <strong>in</strong>sights <strong>in</strong>to whether the basic dynamic commodity price <strong>and</strong> cont<strong>in</strong>gent<br />

claims pric<strong>in</strong>g model is robust over time, we have estimated the model for two equally<br />

long non-overlapp<strong>in</strong>g sub-periods. The relevant sub-samples thus cover the periods: October<br />

1984 to November 1991 <strong>and</strong> November 1991 to March 1999. Rather than purely quantitative<br />

econometric considerations with respect to diagnostic check<strong>in</strong>g of the estimated model, this<br />

<strong>in</strong>vestigation is especially motivated by qualitative considerations such as: Does seasonal patterns<br />

change systematical over time? And, are correlation structures qualitatively robust over<br />

time?<br />

Seasonal patterns may vary over time due to especially production <strong>and</strong> storage <strong>in</strong>novations<br />

24


as well as changes <strong>in</strong> government policies with respect to <strong>in</strong>tervention <strong>in</strong> gra<strong>in</strong> markets. For<br />

example, Frechette (1997) analyzes the “flatter<strong>in</strong>g” of the soybean futures profile up through<br />

the 1970s where the Brazilian soybean production <strong>in</strong>creased rapidly <strong>and</strong> changed the seasonal<br />

pattern <strong>in</strong> world soybean production. However, such dramatic changes <strong>in</strong> world soybean<br />

production patterns have not occurred <strong>in</strong> the time periods relevant for our <strong>in</strong>vestigation. Also,<br />

although US government policies prevail<strong>in</strong>g <strong>in</strong> the 1970s <strong>and</strong> 1980s provided a floor under<br />

gra<strong>in</strong> prices, soybeans had a relatively low support level <strong>and</strong>, thus, also no dramatic changes <strong>in</strong><br />

government <strong>in</strong>tervention policies with respect to soybeans seem to have occurred <strong>in</strong> the relevant<br />

estimation periods. In sum, it is not obvious on beforeh<strong>and</strong> what k<strong>in</strong>d of parameter <strong>in</strong>stability<br />

one would expect to f<strong>in</strong>d due to changes <strong>in</strong> soybean markets conditions when splitt<strong>in</strong>g the data<br />

<strong>in</strong>to sub-samples.<br />

The sub-period estimation results are tabulated <strong>in</strong> Table 4 <strong>and</strong> Table 5, respectively.<br />

[ INSERT TABLE 4 AND TABLE 5 ABOUT HERE ]<br />

We will be brief <strong>in</strong> our discussion of the differences between the estimates <strong>in</strong> the two subperiods<br />

<strong>and</strong> will ma<strong>in</strong>ly po<strong>in</strong>t out <strong>and</strong> focus on differences <strong>in</strong> the persistence of shocks, seasonal<br />

patterns, <strong>and</strong> correlations of relevance to the above full-sample discussion of the relation to<br />

the theory of storage <strong>and</strong> the “<strong>in</strong>verse leverage effect.”<br />

It may first be noted that volatility shocks <strong>and</strong> convenience yield shocks seem<strong>in</strong>gly are more<br />

persistent <strong>in</strong> the first sub-period than the second sub-period, as reflected <strong>in</strong> the estimates of<br />

the mean-reversion parameters κ <strong>and</strong> β. Thus, the estimates of κ <strong>in</strong>dicate that, under the<br />

risk-neutral measure, the half-lives of shocks to volatility equal 0.940 years <strong>in</strong> the first subperiod<br />

<strong>and</strong> 0.160 years <strong>in</strong> the second sub-period (while the full-sample analog is 0.305 years,<br />

as discussed earlier). While this is true under the risk-neutral measure, as reflected <strong>in</strong> options<br />

<strong>and</strong> futures prices, the relevant mean-reversion parameter under the true probability measure<br />

is given by κ − λ v σ v . Hence, under the true probability measure the persistence of shocks to<br />

the volatility is much more similar across the two sub-periods, <strong>and</strong> relevant half-lives are 0.141<br />

years <strong>and</strong> 0.093 years, respectively, as opposed to 0.137 years for the full-sample estimation<br />

results.<br />

In order to see more clearly the estimated seasonality patterns <strong>in</strong> volatilities across the two<br />

sub-periods, we have graphed the relevant seasonal components <strong>in</strong> Figure 4.<br />

[ INSERT FIGURE 4 ABOUT HERE ]<br />

25


The figure shows the estimated seasonal function ν(t) for the full-sample as well as the estimated<br />

seasonal function for the first sub-period, ν 1 (t), <strong>and</strong> the estimated seasonal function<br />

for the second sub-period, ν 2 (t). The estimated seasonal functions seem very similar with<br />

amplitudes of similar magnitude <strong>and</strong> with m<strong>in</strong>ima <strong>and</strong> maxima atta<strong>in</strong>ed at about the same<br />

time of the year. The amplitude is slightly higher <strong>in</strong> the second sub-period <strong>and</strong>, if not simply<br />

due to r<strong>and</strong>omness, this could be due to the above-mentioned price floors ma<strong>in</strong>ta<strong>in</strong>ed by the<br />

US government programs up through the 1980s.<br />

I Figure 5 we have similarly illustrated the seasonal patterns of the mean-revert<strong>in</strong>g level of<br />

the convenience yield as estimated <strong>in</strong> the first sub-period, α 1 (t), the second sub-period, α 2 (t),<br />

as well as for the full sample, α(t).<br />

[ INSERT FIGURE 5 ABOUT HERE ]<br />

While the estimated functions atta<strong>in</strong> m<strong>in</strong>ima <strong>and</strong> maxima at the same times of the year,<br />

the amplitudes seem very different. However, though the amplitude for the first sub-period<br />

estimation is much larger, it may be noted that the mean-reversion tendency is at the same<br />

time much lower than for the second sub-period, as reflected <strong>in</strong> the β estimates. Hence, the<br />

overall seasonal effect on the level of the convenience yield may not differ very substantially<br />

across the two sub-periods.<br />

F<strong>in</strong>ally, it is seen from Table 4 <strong>and</strong> Table 5 that the estimated correlation coefficients (ρ 12 ,<br />

ρ 13 , <strong>and</strong> ρ 23 ) are very similar across the two sub-periods. Thus, the conclusion that, e.g.,<br />

convenience yields do not behave like tim<strong>in</strong>g options is robust over time s<strong>in</strong>ce <strong>in</strong> both subperiods<br />

the correlation coefficient, ρ 23 , between the convenience yield <strong>and</strong> volatility is close to<br />

zero <strong>and</strong>, <strong>in</strong> fact, consistently estimated slightly negative. Also, the “<strong>in</strong>verse leverage effect” is<br />

robust over time s<strong>in</strong>ce the correlation coefficient, ρ 13 , between the spot price <strong>and</strong> the volatility<br />

is consistently estimated positive <strong>and</strong> of the same magnitude across the sub-periods.<br />

4 Conclusions<br />

In the paper we have set up a stochastic volatility model of commodity price behavior <strong>and</strong><br />

estimated the model parameters us<strong>in</strong>g panel-data of both soybean futures <strong>and</strong> soybean options.<br />

The estimation approach thus use the <strong>in</strong>formation <strong>in</strong> both time series characteristics of the<br />

evolvement of soybean product prices as well as the <strong>in</strong>formation <strong>in</strong>herent <strong>in</strong> futures price<br />

structures <strong>and</strong> option price structures, such as implied volatility surfaces.<br />

26


Besides estimat<strong>in</strong>g model parameters, which among other th<strong>in</strong>gs facilitates implementation<br />

of hedg<strong>in</strong>g strategies <strong>and</strong> related applications of the specific model, we f<strong>in</strong>d a clear seasonal<br />

pattern <strong>in</strong> both volatilities <strong>and</strong> convenience yields. Also, our results are to some extent consistent<br />

with the theory of storage <strong>in</strong> the sense that the estimations suggest that convenience yields<br />

tend to be low when <strong>in</strong>ventories/supply is high, <strong>and</strong> vice versa. However, our results do not<br />

support the view that convenience yields are like tim<strong>in</strong>g options s<strong>in</strong>ce convenience yields are<br />

not positively correlated with volatility. Furthermore, we f<strong>in</strong>d a positive correlation between<br />

spot prices <strong>and</strong> volatility, as opposed to the so-called leverage effect <strong>in</strong> equity markets, <strong>and</strong><br />

this is consistent with a different slop<strong>in</strong>g “smirk” <strong>in</strong> the implied volatility curve for soybean<br />

options.<br />

27


References<br />

Backus, D.; Foresi, S., & Telmer, C. (1996). “Aff<strong>in</strong>e Models of Currency Prices”. Work<strong>in</strong>g<br />

paper, New York University.<br />

Bakshi, G. S.; Cao, C., & Chen, Z. (1997). “Empirical Performance of Alternative Option<br />

Pric<strong>in</strong>g Models”. Journal of F<strong>in</strong>ance, 52, 2003–2049.<br />

Bates, D. S. (1996). “Jumps <strong>and</strong> <strong>Stochastic</strong> <strong>Volatility</strong>: Exchange Rate Processes Implicit <strong>in</strong><br />

PHLX Deutsche Mark Options”. Review of F<strong>in</strong>ancial Studies, 9(1), 69–107.<br />

Bates, D. S. (2000). “Post-’87 Crash Fears <strong>in</strong> S&P 500 Futures Option Market”. Journal of<br />

Econometrics, 94(1–2), 181–238.<br />

Bessemb<strong>in</strong>der, H.; Coughenour, J. F.; Segu<strong>in</strong>, P. J., & Smoller, M. M. (1995). “Mean Reversion<br />

<strong>in</strong> Equilibrium Asset Prices: Evidence from the Futures Term Structure”. Journal of<br />

F<strong>in</strong>ance, 50(1), 361–375.<br />

Bessemb<strong>in</strong>der, H.; Coughenour, J. F.; Segu<strong>in</strong>, P. J., & Smoller, M. M. (1996). “Is there a<br />

term structure of futures volatilities: Reevaluat<strong>in</strong>g the Samuelson hypothesis”. Journal<br />

of Fixed Income, 4, 45–58.<br />

Bjerksund, P. (1991). “Cont<strong>in</strong>gent Claims Evaluation when the Convenience Yield is <strong>Stochastic</strong>”.<br />

Work<strong>in</strong>g paper, Norwegian School of Economics <strong>and</strong> Bus<strong>in</strong>ess Adm<strong>in</strong>istration.<br />

Black, F. (1976). “The Pric<strong>in</strong>g of Commodity Contracts”. Journal of F<strong>in</strong>ancial Economics,<br />

3, 167–179.<br />

Brennan, M. J. (1958). “The Supply of Storage”. American Economic Review, 48, 50–72.<br />

Brennan, M. J. (1991). “The Price of Convenience <strong>and</strong> the Valuation of Commodity Cont<strong>in</strong>gent<br />

Claims”. In Lund, D. & Øksendal, B., editors, <strong>Stochastic</strong> Models <strong>and</strong> Option Values, pp.<br />

33–71. Elsevier Science Publishers B.V. (North-Holl<strong>and</strong>).<br />

Brennan, M. J. & Schwartz, E. S. (1985). “Evaluat<strong>in</strong>g Natural Resource Investments”. Journal<br />

of Bus<strong>in</strong>ess, 58, 135–157.<br />

Broadie, M. & Detemple, J. (1996). “American option valuation: new bounds, approximations,<br />

<strong>and</strong> a comparison of exist<strong>in</strong>g methods”. Review of F<strong>in</strong>ancial Studies, 9(4), 1211–1250.<br />

28


Chambers, M. & Bailey, R. (1996). “A Theory of Commodity Price Fluctuations”. Journal of<br />

Political Economy, 104, 924–957.<br />

Chen, R.-R. & Scott, L. (1993). “Maximum Likelihood Estimation for a Multifactor Equilibrium<br />

Model of the Term Structure of Interest Rates”. Journal of Fixed Income, 3(3),<br />

14–31.<br />

Chernov, M. & Ghysels, E. (2000). “A study towards a unified approach to the jo<strong>in</strong>t estimation<br />

of objective <strong>and</strong> risk neutral measures for the purpose of options valuation”. Journal of<br />

F<strong>in</strong>ancial Economics, 56, 407–458.<br />

Cortazar, G. & Schwartz, E. S. (1994). “The Evaluation of Commodity Cont<strong>in</strong>gent Claims”.<br />

Journal of Derivatives, 1, 27–39.<br />

Cox, J. C.; Ingersoll, Jr., J. E., & Ross, S. A. (1981). “The Relation between Forward Prices<br />

<strong>and</strong> Futures Prices”. Journal of F<strong>in</strong>ancial Economics, 9(4), 321–346.<br />

Dai, Q. & S<strong>in</strong>gleton, K. J. (2000). “Specification analysis of aff<strong>in</strong>e term structure models”.<br />

Journal of F<strong>in</strong>ance, 55, 1943–1978.<br />

Deaton, A. & Laroque, G. (1992). “On the Behavior of Commodity Prices”. Review of<br />

Economic Studies, 59, 1–23.<br />

Deaton, A. & Laroque, G. (1996). “Competitive Storage <strong>and</strong> Commodity Price Dynamics”.<br />

Journal of Political Economy, 104, 896–923.<br />

Dixit, A. K. & P<strong>in</strong>dyck, R. S. (1994). Investment under Uncerta<strong>in</strong>ty. Pr<strong>in</strong>ceton University<br />

Press, Pr<strong>in</strong>ceton, New Jersey.<br />

Duffee, G. R. (1999). “Estimat<strong>in</strong>g the price of default risk”. Review of F<strong>in</strong>ancial Studies, 12,<br />

197–226.<br />

Duffie, J. D. (1996). Dynamic Asset Pric<strong>in</strong>g Theory. Pr<strong>in</strong>ceton University Press, Pr<strong>in</strong>ceton,<br />

New Jersey, 2 edition.<br />

Duffie, J. D. & Kan, R. (1996). “A Yield-Factor Model of Interest Rates”. Mathematical<br />

F<strong>in</strong>ance, 6, 379–406.<br />

Duffie, J. D.; Pan, J., & S<strong>in</strong>gleton, K. J. (2000). “Transform analysis <strong>and</strong> asset pric<strong>in</strong>g for<br />

aff<strong>in</strong>e jump-diffusions”. Econometrica, 68, 1343–1376.<br />

29


Duffie, J. D. & S<strong>in</strong>gleton, K. J. (1997). “An Econometric Model of the Term Structure of<br />

Interest Rate Swap Yields”. Journal of F<strong>in</strong>ance, 52, 1287–1321.<br />

Duffie, J. D. & S<strong>in</strong>gleton, K. J. (1999). “Model<strong>in</strong>g term structures of defaultable bonds”.<br />

Review of F<strong>in</strong>ancial Studies, 12, 687–720.<br />

Fama, E. F. & French, K. R. (1987). “Commodity Futures Prices: Some Evidence on Forecast<br />

Power, Premiums, <strong>and</strong> the Theory of Storage”. Journal of Bus<strong>in</strong>ess, 60(1), 55–73.<br />

Frechette, D. L. (1997). “The Dynamics of Convenience <strong>and</strong> the Brazilian Soybean Boom”.<br />

American Journal of Agricultural Economics, 79, 1108–1118.<br />

Gibson, R. & Schwartz, E. S. (1991). “Valuation of Long Term Oil-L<strong>in</strong>ked Assets”. In Lund,<br />

D. & Øksendal, B., editors, <strong>Stochastic</strong> Models <strong>and</strong> Option Values, pp. 73–101. Elsevier<br />

Science Publishers B.V. (North-Holl<strong>and</strong>).<br />

Hamilton, J. D. (1994). Time Series Analysis. Pr<strong>in</strong>ceton University Press, Pr<strong>in</strong>ceton, New<br />

Jersey.<br />

Hannan, E. J.; Terrell, R. D., & Tuckwell, N. (1970). “The Seasonal Adjustment of Economic<br />

Time Series”. International Economic Review, 11, 24–52.<br />

Heston, S. L. (1993). “A closed-Form Solution for Options with <strong>Stochastic</strong> <strong>Volatility</strong> with<br />

Applications to Bond <strong>and</strong> Currency Options”. Review of F<strong>in</strong>ancial Studies, 6(2), 327–<br />

343.<br />

Hilliard, J. E. & Reis, J. (1998). “Valuation of commodity futures <strong>and</strong> options under stochastic<br />

convenience yields, <strong>in</strong>terest rates, <strong>and</strong> jump diffusions <strong>in</strong> the spot”. Journal of F<strong>in</strong>ancial<br />

<strong>and</strong> Quantitative Analysis, 33(1), 61–86.<br />

Hilliard, J. E. & Reis, J. (1999). “Jump processes <strong>in</strong> commodity futures prices <strong>and</strong> options<br />

pric<strong>in</strong>g”. American Journal of Agricultural Economics, 81, 273–286.<br />

Hull, J. C. (2000). Options, Futures, <strong>and</strong> other Derivatives. Prentice-Hall, 4 edition.<br />

Jamshidian, F. & Fe<strong>in</strong>, M. (1990). “Closed-form solutions for oil futures <strong>and</strong> European options<br />

<strong>in</strong> the Gibson-Schwartz model: A note”. Work<strong>in</strong>g paper, Merrill Lynch Capital Markets.<br />

Jarrow, R. A. (1996). Modell<strong>in</strong>g Fixed Income Securities <strong>and</strong> Interest Rate Options. McGraw-<br />

Hill.<br />

30


Kaldor, N. (1939). “Speculation <strong>and</strong> Economic Stability”. Review of Economic Studies, 7,<br />

1–27.<br />

Koekebakker, S. & Lien, G. (2001). “Term structure of volatility <strong>and</strong> price jumps <strong>in</strong> agricultural<br />

markets – Evidence from option data”. Work<strong>in</strong>g paper, Norwegian Agricultural Economics<br />

Research Institute.<br />

McDonald, R. & Siegel, D. (1984). “Option pric<strong>in</strong>g when the underly<strong>in</strong>g asset earns a belowequilibrium<br />

rate of return”. Journal of F<strong>in</strong>ance, 39, 261–265.<br />

McDonald, R. & Siegel, D. (1986). “The value of wait<strong>in</strong>g to <strong>in</strong>vest”. The Quarterly Journal<br />

of Economics, 101, 707–728.<br />

Miltersen, K. R. & Schwartz, E. S. (1998). “Pric<strong>in</strong>g of Options on Commodity Futures with<br />

<strong>Stochastic</strong> Term Structures of Convenience Yields <strong>and</strong> Interest Rates”. Journal of F<strong>in</strong>ancial<br />

<strong>and</strong> Quantitative Analysis, 33(1), 33–59.<br />

Pan, J. (2001). “The jump-risk premia implicit <strong>in</strong> options: Evidence from an <strong>in</strong>tegrated<br />

time-series study”. Work<strong>in</strong>g paper, MIT Sloan School of Management.<br />

Routledge, B. R.; Seppi, D. J., & Spatt, C. S. (2000). “Equilibrium Forward Curves for<br />

Commodities”. Journal of F<strong>in</strong>ance, 55(3), 1297–1338.<br />

Rub<strong>in</strong>ste<strong>in</strong>, M. (1994). “Implied B<strong>in</strong>omial Trees”. Journal of F<strong>in</strong>ance, 49, 771–818.<br />

Samuelson, P. A. (1965). “Proof that properly anticipated prices fluctuate R<strong>and</strong>omly”. Industrial<br />

Management Review, 6, 41–49.<br />

Schwartz, E. S. (1997). “The <strong>Stochastic</strong> Behavior of Commodity Prices: Implications for<br />

Valuation <strong>and</strong> Hedg<strong>in</strong>g”. Journal of F<strong>in</strong>ance, 52(3), 923–973.<br />

Schwartz, E. S. & Smith, J. E. (2000). “Short-Term Variations <strong>and</strong> Long-Term Dynamics <strong>in</strong><br />

Commodity Prices”. Management Science, 46(7), 893–911.<br />

S<strong>in</strong>gleton, K. J. (2001). “Estimation of aff<strong>in</strong>e asset pric<strong>in</strong>g models us<strong>in</strong>g the empirical characteristic<br />

function”. Journal of Econometrics, 102, 111–141.<br />

Sørensen, C. (2001). “Model<strong>in</strong>g seasonality <strong>in</strong> agricultural commodity futures”. Work<strong>in</strong>g<br />

paper, Department of F<strong>in</strong>ance, Copenhagen Bus<strong>in</strong>ess School, Denmark. Forthcom<strong>in</strong>g <strong>in</strong><br />

The Journal of Futures Markets.<br />

31


Telser, L. G. (1958). “Futures Trad<strong>in</strong>g <strong>and</strong> the Storage of Cotton <strong>and</strong> Wheat”. Journal of<br />

Political Economy, 66, 233–255.<br />

White, H. (1982). “Maximum likelihood estimation of misspecified models”. Econometrica,<br />

50, 1–25.<br />

Williams, J. & Wright, B. D. (1991). Storage <strong>and</strong> Commodity Markets. Cambridge University<br />

Press, Cambridge, Engl<strong>and</strong>.<br />

Work<strong>in</strong>g, H. (1948). “Theory of the Inverse Carry<strong>in</strong>g Charge <strong>in</strong> Futures Markets”. Journal of<br />

Farm Economics, 30, 1–28.<br />

Work<strong>in</strong>g, H. (1949). “The Theory of the Price of Storage”. American Economic Review, 39,<br />

1254–1262.<br />

Wright, B. D. & Williams, J. (1989). “A Theory of Negative Prices for Storage”. Journal of<br />

Futures Markets, 9, 1–13.<br />

32


Appendix<br />

Lemma 1 Let p t = log P t . The conditional means <strong>and</strong> conditional variances of the process<br />

X t = (p t , δ t , v t ), described by the stochastic differential equation system (1), (2), <strong>and</strong> (3), are<br />

given by<br />

where<br />

(<br />

E t [X t+∆ ] = Φ(t + ∆) X t +<br />

∫ t+∆<br />

t<br />

)<br />

Φ −1 (u)l(u) du<br />

(∫ t+∆<br />

)<br />

Var t [X t+∆ ] = Φ(t + ∆) E t [v u ] Φ −1 (u)Σ(u)Φ −1 (u) ′ du Φ(t + ∆) ′ (31)<br />

t<br />

⎛ ⎞ ⎛<br />

⎞<br />

r<br />

e ν(u) ρ 12 σ δ e ν(u) ρ 13 σ v e 1 2 ν(u)<br />

l(u) = ⎜ α(u) ⎟<br />

⎝ ⎠ , Σ(u) = ⎜ ρ<br />

⎝ 12 σ δ e ν(u) σδ 2eν(u) ρ 23 σ v σ δ e 1 2 ν(u)<br />

⎟<br />

⎠<br />

θ<br />

ρ 13 σ v e 1 2 ν(u) ρ 23 σ v σ δ e 1 2 ν(u) σv<br />

2<br />

<strong>and</strong> where Φ is the unique solution to the matrix differential equation<br />

with<br />

dΦ(u)<br />

du<br />

(30)<br />

= L(u)Φ(u) , Φ(t) = I (32)<br />

⎛<br />

⎞<br />

0 −1 − 1 2 + λ P e 1 2 ν(u)<br />

L(u) = ⎜ 0 −β λ<br />

⎝<br />

δ σ δ e 1 2 ν(u)<br />

⎟<br />

⎠ .<br />

0 0 −κ + λ v σ v<br />

In particular, the conditional expectations enter<strong>in</strong>g the right h<strong>and</strong> side <strong>in</strong>tegral of (31) are<br />

described by (30) <strong>and</strong> given by E t [v u ] = v t e −κ(u−t) + θ κ (1 − e−κ(u−t) ).<br />

Proof:<br />

The dynamics of X t = (p t , δ t , v t ) can be written <strong>in</strong> matrix form as<br />

dX t = (L(t)X t + l(t)) dt + √ v t σ(t) dW t (33)<br />

where σ(t) = diag[e 1 2 ν(t) , σ δ e 1 2 ν(t) , σ v ] is a diagonal matrix with the three arguments <strong>in</strong> the<br />

diagonal <strong>and</strong> zeros off the diagonal. By us<strong>in</strong>g Ito’s lemma, <strong>and</strong> the def<strong>in</strong>ition of Φ <strong>in</strong> (32), it<br />

is seen that X can be characterized as the solution to (s ≥ t)<br />

( ∫ s<br />

∫ s<br />

)<br />

X s = Φ(s) X t + Φ −1 √<br />

(u)l(u) du + vu Φ −1 (u)σ(u) dW u . (34)<br />

Note that ∫ s<br />

t<br />

t<br />

t<br />

√<br />

vu Φ −1 (u)σ(u) dW u , s ≥ t, is a mart<strong>in</strong>gale start<strong>in</strong>g at zero at time t. Hence,<br />

by tak<strong>in</strong>g conditional expectations <strong>in</strong> (34) (<strong>and</strong> sett<strong>in</strong>g s = t + ∆), one obta<strong>in</strong>s (30). Furthermore,<br />

s<strong>in</strong>ce the two first terms on the right h<strong>and</strong> side of (34) are F t -measurable, the result<br />

<strong>in</strong> (31) follows by an application of e.g. Karatzas <strong>and</strong> Shreve (1991), Proposition 2.17, p. 144.<br />

33


Table 1: Summary statistics for soybean futures, 1984-1999.<br />

Contracts<br />

Days to maturity<br />

Settlement prices<br />

Number of<br />

observations Mean Mean Std. deviation<br />

All 6024 209.53 627.49 86.08<br />

Grouped <strong>in</strong>to time to maturity:<br />

1. Maturity 753 28.60 622.86 97.13<br />

2. Maturity 753 81.75 624.82 95.24<br />

3. Maturity 753 132.52 626.84 92.24<br />

4. Maturity 753 183.57 627.88 87.92<br />

5. Maturity 753 234.00 628.75 83.81<br />

6. Maturity 753 284.36 628.99 80.09<br />

7. Maturity 753 337.77 629.18 76.38<br />

8. Maturity 753 393.63 630.59 72.68<br />

Grouped <strong>in</strong>to expiration months:<br />

January 893 215.46 622.51 81.96<br />

March 877 210.84 628.02 85.24<br />

May 872 212.17 634.03 87.13<br />

July 883 214.26 638.37 91.18<br />

August 812 197.62 632.63 90.16<br />

September 808 197.84 621.02 84.49<br />

November 879 216.55 615.77 80.09<br />

34


Table 2: Summary statistics for soybean call options, 1984-1999.<br />

Contracts<br />

Days to maturity<br />

Implied volatilities<br />

Number of<br />

observations Mean Mean Std. deviation<br />

All 3765 71.95 19.99 6.60<br />

Grouped <strong>in</strong>to time to maturity <strong>and</strong> moneyness:<br />

1. Maturity – ITM ∗ 753 40.50 18.88 6.53<br />

1. Maturity – ATM ∗ 753 40.50 19.61 6.99<br />

1. Maturity – OTM ∗ 753 40.50 21.58 7.19<br />

2. Maturity – ATM ∗ 753 93.71 19.81 6.38<br />

3. Maturity – ATM ∗ 753 144.69 20.06 5.51<br />

Grouped <strong>in</strong>to expiration months:<br />

January 548 66.65 18.13 4.72<br />

March 636 77.32 17.13 3.72<br />

May 645 78.49 16.71 4.17<br />

July 631 81.50 19.40 4.84<br />

August 431 71.85 25.17 8.40<br />

September 374 59.04 26.28 8.76<br />

November 500 60.17 21.47 5.74<br />

∗ The average call prices are: 30.58 for 1. Maturity – ITM, 15.68 for 1. Maturity – ATM, 8.31<br />

for 1. Maturity – OTM, 25.25 for 2. Maturity – ATM, <strong>and</strong> 31.37 for 3. Maturity – ITM.<br />

35


Table 3: Parameter estimates for soybean contracts, Full sample: Oct. 1984 – Mar. 1999.<br />

<strong>Volatility</strong> κ: 2.2708 θ : 0.1542 σ v : 0.7414 ν 1 : −0.4096 ν ∗ 1 : −0.2626 ν 2 : 0.2250 ν ∗ 2 : 0.1293<br />

parameters: (0.2654) (0.0085) (0.0299) (0.0387) (0.0483) (0.0229) (0.0223)<br />

Convenience yield β : 0.8145 σ δ : 0.9933 α 0 : 0.0612 α 1 : −0.2683 α ∗ 1 : 0.1915 α 2 : 0.2338 α ∗ 2 : 0.2689<br />

parameters: (0.2100) (0.1345) (0.0086) (0.0276) (0.0130) (0.0255) (0.0600)<br />

Correlations <strong>and</strong> risk ρ 12 : 0.3979 ρ 13 : 0.4078 ρ 23 : −0.0784 λ P : −0.6274 λ δ : −0.1282 λ v : −3.7829<br />

premia parameters: (0.0110) (0.0461) (0.0772) (1.3001) (1.2286) (1.5610)<br />

36<br />

Correlation matrix of measurement errors (st<strong>and</strong>ard deviation estimates <strong>in</strong> diagonal):<br />

⎛<br />

⎜<br />

⎝<br />

⎞<br />

0.9606<br />

−0.3075 0.8040<br />

0.0489 0.2150 2.4068<br />

0.0164 0.1373 0.7310 3.4679<br />

−0.2480 0.1719 0.1118 0.1053 0.0180<br />

−0.1871 0.1153 0.1435 0.1549 0.8947 0.0265<br />

−0.1236 0.0850 0.2187 0.2398 0.7175 0.9137 0.0279<br />

−0.0638 0.0566 0.2922 0.3084 0.5302 0.7388 0.9238 0.0254<br />

⎟<br />

−0.0688 0.0755 0.2306 0.2765 0.3838 0.5751 0.7466 0.8870 0.0193 ⎠<br />

−0.1199 0.0808 0.1573 0.2042 0.2053 0.3663 0.4892 0.6014 0.8191 0.0116<br />

Log-likelihood function value: 8468.59


Table 4: Parameter estimates for soybean contracts, 1. half: Oct. 1984 – Nov. 1991.<br />

<strong>Volatility</strong> κ: 0.7373 θ : 0.1103 σ v : 0.7626 ν 1 : −0.3825 ν ∗ 1 : −0.1448 ν 2 : 0.2461 ν ∗ 2 : 0.0861<br />

parameters: (0.5038) (0.0110) (0.0282) (0.0371) (0.0423) (0.0332) (0.0186)<br />

Convenience yield β : 0.3186 σ δ : 0.7385 α 0 : 0.0872 α 1 : −0.2103 α ∗ 1 : 0.1571 α 2 : 0.2300 α ∗ 2 : 0.2454<br />

parameters: (0.1921) (0.0887) (0.0090) (0.0336) (0.0139) (0.0335) (0.0516)<br />

Correlations <strong>and</strong> risk ρ 12 : 0.3549 ρ 13 : 0.4989 ρ 23 : −0.0638 λ P : −3.0441 λ δ : −1.9042 λ v : −5.4792<br />

premia parameters: (0.0150) (0.1052) (0.0836) (2.0377) (1.6564) (1.8105)<br />

37<br />

Correlation matrix of measurement errors (st<strong>and</strong>ard deviation estimates <strong>in</strong> diagonal):<br />

⎛<br />

⎜<br />

⎝<br />

⎞<br />

1.0093<br />

−0.1240 0.8056<br />

0.1506 0.2121 2.6630<br />

0.0842 0.1804 0.6474 3.6959<br />

−0.2235 0.1148 0.1393 0.0651 0.0142<br />

−0.1720 0.0662 0.0958 0.0409 0.9347 0.0219<br />

−0.1396 0.0662 0.1087 0.0256 0.8492 0.9599 0.0251<br />

−0.1216 0.0656 0.1563 0.0505 0.7250 0.8476 0.9381 0.0244<br />

⎟<br />

−0.1333 0.0901 0.0660 0.0475 0.5411 0.6404 0.7345 0.8644 0.0206 ⎠<br />

−0.1787 0.0821 −0.0131 0.0266 0.3112 0.3578 0.4195 0.5321 0.8027 0.0129<br />

Log-likelihood function value: 4205.60


Table 5: Parameter estimates for soybean contracts, 2. half: Nov. 1991 – Mar. 1999.<br />

<strong>Volatility</strong> κ: 4.3447 θ : 0.2572 σ v : 0.8736 ν 1 : −0.4185 ν ∗ 1 : −0.2587 ν 2 : 0.2193 ν ∗ 2 : 0.1421<br />

parameters: (0.3286) (0.0144) (0.0351) (0.0190) (0.0268) (0.0108) (0.0103)<br />

Convenience yield β : 1.1901 σ δ : 1.1621 α 0 : 0.0328 α 1 : −0.1875 α ∗ 1 : 0.1777 α 2 : 0.1489 α ∗ 2 : 0.1879<br />

parameters: (0.0794) (0.1401) (0.0047) (0.0068) (0.0073) (0.0039) (0.0035)<br />

Correlations <strong>and</strong> risk ρ 12 : 0.3972 ρ 13 : 0.5085 ρ 23 : −0.0409 λ P : −0.7322 λ δ : −1.8687 λ v : −3.5911<br />

premia parameters: (0.0115) (0.0361) (0.1197) (2.1073) (1.9719) (2.8661)<br />

38<br />

Correlation matrix of measurement errors (st<strong>and</strong>ard deviation estimates <strong>in</strong> diagonal):<br />

⎛<br />

⎜<br />

⎝<br />

⎞<br />

0.9013<br />

−0.4926 0.7990<br />

−0.1575 0.3329 1.9643<br />

−0.1543 0.2437 0.7839 2.5335<br />

−0.2784 0.1928 0.0113 0.0066 0.0213<br />

−0.2068 0.1145 0.0766 0.0743 0.8746 0.0303<br />

−0.1114 0.0506 0.1487 0.1832 0.6274 0.8831 0.0297<br />

−0.0207 0.0071 0.2015 0.2410 0.3607 0.6371 0.8998 0.0254<br />

⎟<br />

−0.0263 0.0403 0.1731 0.1870 0.1877 0.4560 0.6987 0.8859 0.0178 ⎠<br />

−0.0891 0.0948 0.1543 0.1252 0.0098 0.2526 0.4272 0.5733 0.8113 0.0107<br />

Log-likelihood function value: 4642.37


1100.00<br />

Futures prices <strong>in</strong> cents per bushel<br />

1000.00<br />

900.00<br />

800.00<br />

700.00<br />

600.00<br />

500.00<br />

400.00<br />

Jan-84 Jan-88 Jan-92 Jan-96 Jan-2000<br />

Short maturity<br />

Long maturity<br />

Figure 1: Time series of soybean futures prices. The figure displays the futures prices on<br />

the contracts with the shortest (1. Maturity) time to maturity <strong>and</strong> the longest (8. Maturity)<br />

time to maturity over the full sample period from October 1984 to March 1999.<br />

39


0.80<br />

0.70<br />

0.60<br />

Implied volatilities<br />

0.50<br />

0.40<br />

0.30<br />

0.20<br />

0.10<br />

0.00<br />

Jan-84 Jan-88 Jan-92 Jan-96 Jan-2000<br />

Figure 2: Time series of implied volatilities on soybean call options. The figure<br />

displays the implied volatilities on the at-the-money call option contracts with the shortest (1.<br />

Maturity) time to maturity over the full sample period from October 1984 to March 1999.<br />

40


1.00<br />

0.75<br />

0.50<br />

0.25<br />

0.00<br />

- 0.25<br />

α(t)<br />

ν(t)<br />

frag replacements<br />

- 0.50<br />

- 0.75<br />

Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec<br />

Figure 3: The seasonal functions α(t) <strong>and</strong> ν(t). The figure displays the target seasonal<br />

level of convenience yields, α(t), <strong>and</strong> the seasonal component <strong>in</strong> the soybean price volatility,<br />

ν(t).<br />

41


0.50<br />

0.25<br />

0.00<br />

ν 1 (t)<br />

ν 2 (t)<br />

ν(t)<br />

- 0.25<br />

frag replacements<br />

- 0.50<br />

Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec<br />

Figure 4: The three volatility seasonal functions ν(t), ν 1 (t), <strong>and</strong> ν 2 (t). The figure<br />

displays the seasonal component <strong>in</strong> the soybean price volatility for the full sample, ν(t), the<br />

first sub-period, ν 1 (t), <strong>and</strong> the second sub-period, ν 2 (t)<br />

42


2.00<br />

1.50<br />

1.00<br />

0.50<br />

0.00<br />

- 0.50<br />

α 1 (t)<br />

α(t)<br />

α 2 (t)<br />

frag replacements<br />

- 1.00<br />

- 1.50<br />

Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec<br />

Figure 5: The three convenience yield seasonal functions α(t), α 1 (t), <strong>and</strong> α 2 (t). The<br />

figure displays the relative seasonal level of convenience yields for the full sample, α(t), the<br />

first sub-period, α 1 (t), <strong>and</strong> the second sub-period, α 2 (t)<br />

43

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!